Skip to main content Accessibility help
×
Hostname: page-component-5c6d5d7d68-wbk2r Total loading time: 0 Render date: 2024-08-16T10:40:04.338Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  02 December 2009

Brian McGowran
Affiliation:
University of Adelaide
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Biostratigraphy
Microfossils and Geological Time
, pp. 397 - 446
Publisher: Cambridge University Press
Print publication year: 2005

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

Abele, C., Gloe, C. S., Hocking, J. B. et al. (1976). Tertiary. In Geology of Victoria, ed. Douglas, J. G. & Ferguson, J. G., pp. 177–274. Melbourne: Geological Society of Australia Special Publication 5.Google Scholar
Abreu, V. S. & Haddad, G. A. (1998). Glacioeustatic fluctuations: the mechanism linking stable isotope events and sequence stratigraphy from the Early Oligocene to middle Miocene. In Mesozoic and Cenozoic Sequence Stratigraphy of European Basins, ed. Graciansky, P.-C., Hardenbol, J., Jacquin, T. & Vail, P. R., Society of Sedimentary Geology Special Publication No. 60, 245–60.CrossRefGoogle Scholar
Abreu, V. S., Hardenbol, J., Haddad, G. A. et al. (1998). Oxygen isotope synthesis: a Cretaceous ice-house? In Mesozoic and Cenozoic Sequence Stratigraphy of European Basins, ed. Graciansky, P.-C., Hardenbol, J., Jacquin, T. & Vail, P. R.. Society of Sedimentary Geology, Special Publication No. 60, 75–80.CrossRefGoogle Scholar
Adams, C. G. (1965). The foraminifera and stratigraphy of the Melinau Limestone, Sarawak, and its importance in Tertiary correlation. Quarterly Journal of the Geological Society of London, 121, 283–338.CrossRefGoogle Scholar
Adams, C. G.(1967). Tertiary foraminifera in the Tethys, American and Indo-Pacific provinces. Systematics Association Special Publication, 7, 195–217.Google Scholar
Adams, C. G.(1970). A reconsideration of the East Indies Letter Classification of the Tertiary. Bulletin British Museum (Natural History), London, 19, 87–137.Google Scholar
Adams, C. G.(1983). Speciation, phylogenesis, tectonism, climate and eustasy: factors in the evolution of Cenozoic larger foraminiferal bioprovinces. In Evolution Time and Space: The Emergence of the Biosphere, edited Sims, R. W., Price, J. H. & Whalley, P. E. S.. London: Academic Press, Systematics Association Special Volume 23, 255–289.Google Scholar
Adams, C. G.(1984). Neogene larger foraminifera, evolutionary and geological events in the context of datum planes. In Pacific Neogene Datum Planes, ed. Ikebe, N. & Tsuchi, R., pp. 47–67. Tokyo: University of Tokyo Press.Google Scholar
Adams, C. G.(1992). Larger foraminifera and the dating of Neogene events. In Pacific Neogene: Environment, Evolution, and Events, ed. Tsuchi, R. & Ingle, J. C, pp. 221–235. Tokyo: University of Tokyo Press.Google Scholar
Adams, C. G., Butterlin, J. & Samanta, B. K. (1986). Larger foraminifera and events at the Eocene–Oligocene boundary in the Indo-West Pacific region. In Terminal Eocene Events, ed. Pomerol, Ch. & Silva, I. Premoli, pp. 237–252. New York: Elsevier Science.Google Scholar
Adams, C. G., Lee, D E. & Rosen, B. R. (1990). Conflicting isotopic and biotic evidence for tropical sea-surface temperatures during the Tertiary. Palaeogeography, Palaeoclimatology, Palaeoecology, 77, 289–313.CrossRefGoogle Scholar
Ager, D. V. (1984). The stratigraphical code and what it implies. In Catastrophes and Earth History, ed. Berggren, W. A. & Couvering, J. A., pp. 91–100. Princeton: Princeton University Press.CrossRefGoogle Scholar
Agterberg, F. P. & Gradstein, F. M. (1999). The RASC method for ranking and scaling of biostratigraphic events. Earth-Science Reviews, 46, 1–25.CrossRefGoogle Scholar
Allan, R. S. (1933). On the System and Stage names applied to subdivisions of the Tertiary Strata of New Zealand. Transactions New Zealand Institute. 63, 81–108.Google Scholar
Allan, R. S.(1966). The unity of stratigraphy. New Zealand Journal of Geology & Geophysics, 9, 491–494.CrossRefGoogle Scholar
Allmon, W. D. & Bottjer, D. J. (2001). Evolutionary Paleoecology: The Ecological Context of Macroevolutionary Change. New York: Columbia University Press.CrossRefGoogle Scholar
Alroy, J. (1992). Conjunction among taxonomic distributions and the Miocene mammalian biochronology of the Great Plains. Paleobiology, 18, 326–343.CrossRefGoogle Scholar
Alroy, J.(1994). Appearance event ordination: a new biochronological method. Paleobiology, 20, 191–207.CrossRefGoogle Scholar
Alroy, J.(1998a). Diachrony of mammalian appearance events: implications for biochronology. Geology, 26, 23–27.2.3.CO;2>CrossRefGoogle Scholar
Alroy, J.(1998b). Equilibrial diversity dynamics in North American mammals. In Biodiversity Dynamics: Turnover of Populations, Taxa and Communities, ed. McKinney, M. L. & Drake, J., Columbia UP, pp. 232–287.Google Scholar
Alroy, J.(2000). New methods for quantifying macroevolutionary patterns and processes. Paleobiology, 26, 707–733.2.0.CO;2>CrossRefGoogle Scholar
Alroy, J., Koch, P. L. & Zachos, J. C. (2000). In Deep Time: Paleobiology's Perspective, ed. D. H. Erwin & S. L. Wing, Palaeobiology, Supplement to vol. 26(4), 259–288.
Alvarez, W., Arthur, M. A., Fischer, A. G.et al. (1977). Upper Cretaceous–Paleocene magnetic stratigraphy at Gubbio, Italy. V. Type section for the Late Cretaceous–Paleocene geomagnetic reversal scale. Geological Society of America, Bulletin, 88, 383–389.2.0.CO;2>CrossRefGoogle Scholar
Alvarez, L. W., Alvarez, W., Asaro, F. & Michel, H. V. (1980). Extraterrestrial cause for the Cretaceous–Tertiary extinction. Science, 208, 1095–1118.CrossRefGoogle ScholarPubMed
Andreasen, D. J. & Ravelo, A. C. (1997). Tropical Pacific Ocean thermocline depth reconstruction for the last glacial maximum. Paleoceanography, 12, 395–413.CrossRefGoogle Scholar
Applin, E. R., Ellisor, A. E. & Kniker, H. T. (1925). Subsurface stratigraphy of the coastal plain of Texas and Louisiana. American Association of Petroleum Geologists, Bulletin, 9, 79–122.Google Scholar
Arkell, W. J. (1933). The Jurassic System in Great Britain. Oxford: Clarendon Press.Google Scholar
Arkell, W. J.(1956). Comments on stratigraphic procedure and terminology. American Journal of Science, 254, 457–467.CrossRefGoogle Scholar
Armentrout, J. M. (1996). High resolution sequence biostratigraphy: examples from the Gulf of Mexico Plio-Pleistocene. In High Resolution Sequence Stratigraphy: Innovations and Applications, ed. Howell, J. A. & Aitken, J. F., pp. 65–86. The Geological Society Special Publication No. 104.Google Scholar
Arnold, A. J. (1983). Phyletic evalution in the Globorotalia crassaformis (Galloway and Wissler) lineage: a preliminary report. Palaeobiology, 9, 390–8.CrossRefGoogle Scholar
Arnold, A. J. & Parker, W. C. (1999). Biogeography of planktonic foraminifera. In Modern Foraminifera, ed. Gupta, B. K. Sen, pp. 103–122. Dordrecht: Kluwer Academic Publishers.Google Scholar
Aubry, M.-P. (1995). From chronology to stratigraphy: interpreting the lower and middle Eocene stratigraphic record in the Atlantic Ocean. In Geochronology Time Scales and Global Stratigraphic Correlation, ed. Berggren, W. A., Kent, D. V., Aubry, M.-P. & Hardenbol, J.. Tulsa, SEPM Special Publication54, 213–274.CrossRefGoogle Scholar
Aubry, M.-P.(2000). Where should the Global Stratotype Section and Point (GSSP) for the Paleocene–Eocene boundary be located? Comment définir la Limite Paléocène–Eocène? Bulletin de la Societé géologique de France, 171.Google Scholar
Aubry, M.-P. & Berggren, W. A. (2000). The homeless GSSP: the dilemma of the Paleocene/Eocene boundary. Tertiary Research, 20, 107–112.Google Scholar
Aubry, M.-P., Berggren, W. A., Kent, D. V.et al. (1988). Paleogene geochronology: an integrated approach. Paleoceanography, 3, 707–742.CrossRefGoogle Scholar
Aubry, M.-P., Berggren, W. A., Stott, L. & Sinha, A. (1996). The upper Paleocene-lower Eocene stratigraphic record and the Paleocene/Eocene boundary carbon isotope excursions. In Correlation of the Early Paleogene in northwest Europe, ed. Knox, R. W. O'B, Corfield, R. M. & Dunay, R. E., p. 353–380. London: Geological Society, Special Publication No. 101.Google Scholar
Aubry, M.-P., Berggren, W. A., Couvering, J. A. & Steininger, F. (1999). Problems in chronostratigraphy: stages, series, unit and boundary stratotypes, global stratotype section and point and tarnished golden spikes. Earth-Science Reviews, 46, 99–148.CrossRefGoogle Scholar
Aubry, M.-P., Cramer, B. S., Miller, K. G., Wright, J. D., Kent, D. V. & Olsson, R. K. (2000). Late Paleocene event chronology: unconformities, not diachrony. Bulletin de la Societé géologique de France, 171, 367–378.CrossRefGoogle Scholar
Aubry, M.-P., Lucas, S. G. & Berggren, W. A., eds (1998). Late Paleocene–Early Eocene Climatic and Biotic Events in the Marine and Terrestrial Records. New York: Columbia University Press, 513 pp.Google Scholar
Australasian Petroleum Company (1961). The geological results of petroleum exploration in western Papua. Journal of the Geological Society of Australia, 8, 1–133.CrossRef
Ayala, F. (1983). Microevolution and macroevolution. In Evolution from Molecules to Men, ed. Bendall, D. S., pp. 387–402. Cambridge: Cambridge University Press.Google Scholar
Backman, J. & Raffi, I. (1997). Calibration of Miocene nannofossil events to orbitally tuned cyclostratigraphies from Ceara Rise. Proceedings of the Ocean Drilling Program, Scientific Results, 154, 83–99. College Station, Texas.Google Scholar
Bakker, R. (1986). The Dinosaur Heresies. London: Longman Scientific & Technical.Google Scholar
Baldi, T. (1980). The early history of the Paratethys. Földtani Közlöny, 110, 1–18.Google Scholar
Bambach, R. K. & Bennington, J. B. (1996). Statistical testing for paleocommunity recurrence: are similar fossil assemblages ever the same? In New perspectives of faunal stability in the fossil record, ed. L. C., Ivany & K. M. Schopf, Palaeogeography, Palaeoclimatology, Palaeoecology, 127, 103–133.
Bandy, O. L. (1972). Origin and development of Globorotalia (Turborotalia) pachyderma (Ehrenberg). Micropaleontology, 18, 294–318.CrossRefGoogle Scholar
Bandy, O. L., Frerichs, W. E. & Vincent, E. (1967). Origin, development and geologic significance of Neogloboquadrina Bandy, Frerichs and Vincent, gen. nov. Contributions Cushman Foundation Foraminiferal Research, 18, 152–157.Google Scholar
Banner, F. T. & Blow, W. H. (1959). The classification and stratigraphical distribution of the Globigerinaceae. Palaeontology, 2, 1–27.Google Scholar
Banner, F. T. & Blow, W. H.(1965). Progress in the planktonic foraminiferal biostratigraphy of the Neogene. Nature, 208, 1164–1166.CrossRefGoogle Scholar
Barbin, V. (1988). In The Eocene–Oligocene Boundary in the Marche–Umbria Basin (Italy), ed. I. Premoli Silva, R. Coccioni & A. Montanari, p. 163–171. International Subcommission on Paleogene Stratigraphy, International Union of Geological Sciences, Eocene–Oligocene meeting, Ancona (Italy), Special Publication.
Barry, J. C., Morgan, M. E., Flynn, L. J.et al. (2002). Faunal and environmental change in the Late Miocene Siwaliks of northern Pakistan. Paleobiology Memoirs, Suppl. to Vol. 28(2), Memoir 3, 1–72.CrossRefGoogle Scholar
Bé, A. W. H. (1977). An ecological, zoogeographic and taxonomic review of recent planktonic foraminifera. In Oceanic Micropalaeontology, ed. Ramsay, A. T. S., pp. 1–100. London: Academic Press.Google Scholar
Bé, A. W. H.(1982). Biology of plankton foraminifera. In Foraminifera: Notes for a Short Course, ed. T. W. Broadhead, Studies in Ecology, 6, 51–92.
Bé, A. W. H. & Tolderlund, D. S. (1971). Distribution and ecology of living planktonic foraminifera in surface waters of the Atlantic and Indian Oceans. In, Micropaleontology of the Oceans, ed. Funnell, B. M. & Riedel, W. R., pp. 105–150. Cambridge: Cambridge University Press.Google Scholar
Behrensmeyer, A. K., Todd, N. E., Potts, R. & McBrinn, G. E. (1997). Late Pliocene faunal turnover in the Turkana Basin, Kenya and Ethiopia. Science, 278, 1589–1594.CrossRefGoogle ScholarPubMed
Bennett, K. D. (1990). Milankovitch cycles and their effects on species in ecological and evolutionary time. Paleobiology, 16, 11–21.CrossRefGoogle Scholar
Bennett, K. D.(1997). Evolution and Ecology: The Pace of Life. Cambridge: Cambridge University Press.Google Scholar
Benson, R. A. (1995). Editorial: is the death of an ocean falling through a stratigraphic crack? Paleoceanography, 10, 1–3.CrossRefGoogle Scholar
Benson, R. A. & Hodell, D. A. (1994). Comment on ‘A critical re-evaluation of the Miocene–Pliocene boundary as defined in the Mediterranean’ by F. J. Hilgen and C. G. Langereis. Earth & Planetary Science Letters, 124, 245–250.CrossRefGoogle Scholar
Benton, M. J. & Pearson, P. N. (2001). Speciation in the fossil record. Trends in Ecology & Evolution, 16, 405–411.CrossRefGoogle ScholarPubMed
Berger, W. H. (1970). Biogenous deep-sea sediments: fractionation by deep-sea circulation. Geological Society of America Bulletin, 81, 1385–1402.CrossRefGoogle Scholar
Berger, W. H.(1974). Deep-sea sedimentation. In The Geology of Continental Margins, ed. Burk, C. A. & Drake, C. L., pp. 213–241. New York: Springer-Verlag.CrossRefGoogle Scholar
Berger, W. H.(1982). Deep-sea stratigraphy: Cenozoic climatic steps and the search for chemo-climatic feedback. In Cyclic and Event Stratification, ed. Einsele, G. & Seilacher, A., pp. 121–157. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Berger, A. & Loutre, M. F. (1991). Insolation values for the climate of the last 10 million years. Quaternary Science Reviews, 10, 297–317.CrossRefGoogle Scholar
Berger, W. H. & Vincent, E. (1981). Chemostratigraphy and biostratigraphic correlation: exercises in systemic stratigraphy. Proceedings 26th International Geological Congress, Oceanologica Acta 4 (Supplement), 115–127.Google Scholar
Berger, W. H. & Vincent, E.(1986). Deep-sea carbonates: reading the carbon-isotope signal. Geologische Rundschau, 75, 249–269.CrossRefGoogle Scholar
Berggren, W. A. (1960). Paleogene biostratigraphy and planktonic foraminifera of the SW Soviet Union: an analysis of recent Soviet publications. Stockholm Contributions in Geology, 6(5), 63–125.Google Scholar
Berggren, W. A.(1964). The Maestrichtian, Danian and Montian Stages and the Cretaceous–Tertiary boundary. Stockholm Contributions in Geology, 11, 103–176.Google Scholar
Berggren, W. A.(1968). Phylogenetic and taxonomic problems of some Tertiary planktonic foraminiferal lineages. Tulane Studies in Geology, 6, 1–22.Google Scholar
Berggren, W. A.(1969a). Cenozoic chronostratigraphy, planktonic foraminiferal zonation and the radiometric time scale. Nature, 224, 1072–1075.CrossRefGoogle Scholar
Berggren, W. A.(1969b). Rates of evolution in some Cenozoic planktonic foraminifera. Micropaleontology, 15, 351–365.CrossRefGoogle Scholar
Berggren, W. A.(1971a). Multiple phylogenetic zonations of the Cenozoic based on planktonic foraminifera. In Proceedings of the II Planktonics Conference, Roma 1970, ed. Farinacci, A., 41–56. Roma: Edizioni Technoscienza.Google Scholar
Berggren, W. A.(1971b). Tertiary boundaries and correlations. In The Micropaleontology of the Oceans, ed. Funnell, B. F. & Riedel, W. R., pp. 693–808. Cambridge: Cambridge University Press.Google Scholar
Berggren, W. A.(1978). Biochronology. In Contributions to the Geological Time Scale, ed. Cohee, G. V., Glaessner, M. F. & Hedberg, H. D.. American Association of Petroleum Geologists, Studies in Geology, 6, 39–55.Google Scholar
Berggren, W. A.(1984). Neogene planktonic foraminiferal biostratigraphy and biogeography: Atlantic, Mediterranean, and Indo-Pacific regions. In Pacific Neogene Datum Planes: Contributions to Biostratigraphy and Chronology, ed. Ikebe, N. & Tsuchi, R., pp. 111–161. University of Tokyo Press.Google Scholar
Berggren, W. A.(1998). The Cenozoic Era: Lyellian (chrono)stratigraphy and nomenclatural reform at the milennium. In Lyell: The Past is the Key to the Present, ed. Blundell, D. J. & Scott, A. C.. Geological Society Special Publication No. 143, 111–132.Google Scholar
Berggren, W. A. & Aubry, M.-P. (1996). A late Paleocene–early Eocene NW European and North Sea magnetobiochronological correlation network. In Correlation of the Early Paleogene in Northwest Europe, ed. Knox, R. W. O'B., Corfield, R. M. & Dunay, R. E., pp. 309–352. London: Geological Society, Special Publication No. 101.Google Scholar
Berggren, W. A. & Miller, K. G. (1988). Paleogene tropical planktonic foraminiferal biostratigraphy and magnetobiochronology. Micropaleontology, 34, 362–380.CrossRefGoogle Scholar
Berggren, W. A. & Norris, R. D. (1997). Biostratigraphy, phylogeny and systematics of Paleocene trochospiral foraminifera. Micropaleontology, 43, Supplement 1, 1–116.CrossRefGoogle Scholar
Berggren, W. A. & Prothero, D. R. (1992). Eocene–Oligocene climatic and biotic evolution: an overview. In Eocene–Oligocene Climatic and Biotic Evolution, ed. Prothero, D. R. & Berggren, W. A., pp. 1–28. Princeton: Princeton University Press.CrossRefGoogle Scholar
Berggren, W. A. & van Couvering, J. A. (1978). Biochronology. In Contributions to the Geological Time Scale, ed. Cohee, G. V., Glaessner, M. F. & Hedberg, H. D.. American Association of Petroleum Geologists, Studies in Geology, 6, 39–55.Google Scholar
Berggren, W. A., Hilgen, F. J., Langereis, C. G.et al. (1995b). Late Neogene chronology: New perspectives in high-resolution stratigraphy. Geological Society of America, Bulletin, 107(11), 1272–1287.2.3.CO;2>CrossRefGoogle Scholar
Berggren, W. A., Kent, D. V. & Flynn, J. J. (1985). Paleogene geochronology and chronostratigraphy. In The Chronology of the Geological Record, ed. Snelling, N. J.. Geological Society of London, Memoir 10, 141–195.Google Scholar
Berggren, W. A., Kent, D. V., Flynn, J. J. & Couvering, J. A. (1985). Cenozoic geochronology. Geological Society of America, Bulletin, 96, 1407–1418.2.0.CO;2>CrossRefGoogle Scholar
Berggren, W. A., Kent, D. V., Swisher, C. C. III & Aubry, M.-P. (1995). A revised Cenozoic geochronology and chronostratigraphy. In Geochronology Time Scales and Global Stratigraphic Correlation, ed. Berggren, W. A., Kent, D. V., Aubry, M.-P. & Hardenbol, J.. Tulsa, SEPM (Society of Sedimentary Geology) Special Publication 54, 129–212.CrossRefGoogle Scholar
Berggren, W. A., Kent, D. V. & Van Couvering, J. A. (1985). Neogene geochronology and chronostratigraphy. In The Chronology of the Geological Record, ed. Snelling, N. J.. Geological Society of London Memoir 10, 211–260.Google Scholar
Berry, W. B. N. (1968). Growth of a Prehistoric Time Scale Based on Organic Evolution. San Francisco: W. H. Freeman & Co.Google Scholar
Berry, W. B. N.(1977). Graptolite biostratigraphy: a wedding of classical principles and current concepts. In Concepts and Methods of Biostratigraphy, ed. Kauffmann, E. G. & Hazel, J. E., pp. 321–338. Stroudsberg, PA: Dowden, Hutchison & Ross, Inc.Google Scholar
Bignot, G. (1985). Elements of Micropalaeontology. London: Graham & Trotman Ltd.Google Scholar
Billups, K., Channell, J. E. T. & Zachos, J. C. (2002). Late Oligocene to early Miocene geochronology and paleoceanography from the subantarctic South Atlantic. Paleoceanography, 17, 4–11.CrossRefGoogle Scholar
Blackwelder, R. E. (1967). Taxonomy: A Text and Reference Book. New York: John Wiley.Google Scholar
Blackwelder, R. E. & Boyden, A. (1952). The nature of systematics. Systematic Zoology, 1, 26–33.CrossRefGoogle Scholar
Blow, W. H. (1956). Origin and evolution of the foraminiferal genus Orbulina d' Orbigny. Micropaleontology, 2, 57–70.CrossRefGoogle Scholar
Blow, W. H.(1959). Age, correlation and biostratigraphy of the Upper Tocuyo (San Lorenzo) and Pozón formations eastern Falcón, Venezuela. Bulletin of American Paleontology, 39, 1–251.Google Scholar
Blow, W. H.(1969). Late middle Eocene to Recent planktonic foraminiferal biostratigraphy. In Proceedings of the First International Conference on Planktonic Microfossils, ed. Brönnimann, P. & Renz, H. H.. Leiden: E. J. Brill, vol. 1, pp. 199–421.Google Scholar
Blow, W. H.(1970). Validity of biostratigraphic correlations based on the Globigerinacea. Micropaleontology, 16, 257–268.CrossRefGoogle Scholar
Blow, W. H.(1979). The Cainozoic Globigerinida: a study of the morphology, taxonomy, evolutionary relationships and the stratigraphical distribution of some Globigerinida (mostly Globigerinacea). 3 volumes: I. text, part I and part 2, section 1, xvii + 752 pp., 118 figures. II. part II, section 2, ix + pp. 753–1413. III. Atlas, xxi pp., 264 pp. Leiden: E. J. Brill.
Blow, W. H. & Banner, F. T. (1962). Part Two: The mid-Tertiary (Upper Eocene to Aquitanian) Globigerinaceae. In, Fundamentals of Mid-Tertiary Stratigraphical Correlation, ed. Eames, F. E., Banner, F. T., Blow, W. H. & Clarke, W. J., pp. 61–151. Cambridge: Cambridge University Press.Google Scholar
Bobe, R. & Eck, G. G. (2001). Responses of African bovids to Pliocene climatic change. Paleobiology Memoirs, 27, Supplement to No. 2, 1–48.2.0.CO;2>CrossRefGoogle Scholar
Boersma, A., Premoli Silva, I. & Hallock, P. (1998). Trophic models for the well-mixed and poorly mixed warm oceans across the Paleocene/Eocene Epoch boundaries. In Late Paleocene–Early Eocene Climatic and Biotic Events in the Marine and Terrestrial Records, ed. Aubry, M.-P., Lucas, S. G. & Berggren, W. A., 204–213. New York: Columbia University Press.Google Scholar
Bolli, H. M. (1950). The direction of coiling in the evolution of some Globorotaliidae. Contributions of the Cushman Foundation of Foraminiferal Research, 1, 82–89.Google Scholar
Bolli, H. M.(1957a). The genera Globigerina and Globorotalia in the Paleocene–lower Eocene Lizard Springs Formation of Trinidad, B. W. I. United States National Museum, Bulletin, 215, 51–81.Google Scholar
Bolli, H. M.(1957b). Planktonic foraminifera from the Eocene Navet and San Fernando Formations of Trinidad, B. W. I. United States National Museum, Bulletin, 215, 155–172.Google Scholar
Bolli, H. M.(1957c). Planktonic foraminifera from the Oligocene–Miocene Cipero and Lengua Formations of Trinidad, B. W. I. United States National Museum, Bulletin, 215, 97–123.Google Scholar
Bolli, H. M.(1966). Zonation of Cretaceous to Pliocene sediments based on planktonic foraminifera. Boletin Informativo, Asociacion Venezolano de Geologia, Mineraria y Petroleo, 9, 3–32.Google Scholar
Bolli, H. M.(1967). The subspecies of Globorotalia fohsi Cushman and Ellisor and the zones based on them. Micropalaeontology, 13, 502–512.CrossRefGoogle Scholar
Bolli, H. M.(1971). The direction of coiling in planktonic foraminifera. In The Micropalaeontology of Oceans, ed. Funnell, B. M. & Riedel, W. R., pp. 639–648. Cambridge: Cambridge University Press.Google Scholar
Bolli, H. M. & Saunders, J. B. (1985). Oligocene to Holocene low-latitude planktic foraminifera. In Plankton Stratigraphy, ed. Bolli, H. M., Saunders, J. B. & Perch-Nielsen, K., pp. 155–262. Cambridge, Cambridge University Press, 2 volumes.Google Scholar
Bolli, H. M., Loeblich, A. R. Jr. & Tappan, H. (1957). Planktonic foraminiferal families Hantkeninidae, Orbulinidae, Globorotaliidae, and Globotruncanidae. United States National Museum, Bulletin, 215, 3–50.Google Scholar
Bolli, H. M., Saunders, J. B. & Perch-Nielsen, K., Editors (1985). Plankton Stratigraphy. Cambridge, Cambridge University Press, 2 volumes.Google Scholar
Bonuso, N., Newton, C. R., Brower, J. C. & Ivany, L. C. (2002). Does coordinated stasis yield taxonomic and ecologic stability?: Middle Devonian Hamilton Group of New York. Geology, 30, 1055–1058.2.0.CO;2>CrossRefGoogle Scholar
Boucot, A. J. (1982). Ecostratigraphy. In, Stratigraphy Quo Vadis? ed. E. Seibold & J. H. Meulenkamp, American Association of Petroleum Geologists, Studies in Geology No. 16, 55–60.
Boucot, A. J.(1983). Does evolution take place in an ecological vacuum? Journal of Paleontology, 57, 1–30.Google Scholar
Boucot, A. J.(1990a). Modern paleontology: using biostratigraphy to the utmost. Revista Española de Paleontología, 5, 63–70.Google Scholar
Boucot, A. J.(1990b). Community evolution: its evolutionary and biostratigraphic significance. In Paleocommunity Temporal Dynamics: The Long-term Development of Multispecies Assemblages, ed. Miller, W. III. The Paleontological Society Special Publication, 5, 48–70.Google Scholar
Boucot, A. J.(1994). The episodic, rather than periodic nature of extinction events. Revista de la Sociedad Mexicana de Paleontologia, 7, 15–35.Google Scholar
Boudagher-Fadel, M. K. (2002). The stratigraphical relationship between planktonic and larger berthic foraminifera in the middle Miocene to lower Pliocene carbonate facies of Sulawesi, Indonesia. Micropalaeontology, 48, 153–76.CrossRefGoogle Scholar
Boudagher-Fadel, M. K. & Bonner, F. T. (1999). Revision of the stratigraphic significance of the Oligocere–Miocere letter Stages. Reveue de Micropaléontologie, 42, 93–7.CrossRefGoogle Scholar
Bralower, T. J., Leckie, R. M., Sliter, W. V. & Thierstein, H. R. (1995). An integrated Cretaceous microfossil biostratigraphy. In Geochronology Time Scales and Global Stratigraphic Correlation, ed. Berggren, W. A., Kent, D. V, Aubry, M.-P. & Hardenbol, J.. SEPM (Society of Sedimentary Geology), Special Publication, 54, 65–80.CrossRefGoogle Scholar
Bramlette, M. N. & Martini, E. (1964). The great change in calcareous nannoplankton fossils between the Maastrichtian and Danian. Micropaleontology, 10, 291–322.CrossRefGoogle Scholar
Brasier, M. D. (1980). Microfossils. London: Allen & Unwin.Google Scholar
Brasier, M. D.(1995). Fossil indicators of nutrient levels. 2. Evolution and extinction in relation to oligotrophy. In Marine Palaeoenvironmental Analysis from Fossils, ed. Bosence, D. J. W. & Allison, P. A.. London: The Geological Society, Special Publication No. 83, 133–150.Google Scholar
Bretsky, P. W. (1979). History of paleontology: post-Darwinian. In Encyclopedia of Paleontology, ed. Fairbridge, R. W. & Jablonski, D., pp. 384–395. Stroudsberg, PA: Dowden, Hutchison & Ross, Inc.Google Scholar
Brett, C. E. (1998). Sequence stratigraphy, paleoecology, and evolution: biotic clues and responses to sea-level fluctuations. Palaios, 13: 241–262.CrossRefGoogle Scholar
Brett, C. E.(2001). A slice of the ‘layer cake’: the paradox of ‘frosting continuity’. Palaios.Google Scholar
Brett, C. E. & Baird, G. C. (1995). Coordinated stasis and evolutionary ecology of Silurian to middle Devonian faunas in the Appalachian Basin. In New Approaches to Speciation in the Fossil Record, ed. Erwin, D. H. & Anstey, R. L.. pp. 285–315, New York: Columbia University Press.Google Scholar
Brett, C. E. & Baird, G. C.(1997). Paleontological Events: Stratigraphic, Ecological, and Evolutionary Implications. New York: Columbia University Press.Google Scholar
Brett, C. E., Ivany, L. C. & Schopf, K. M. (1996). Coordinated stasis: an overview. Palaeogeography, Palaeoclimatology, Palaeoecology, 127, 1–20.CrossRefGoogle Scholar
Brinkhuis, H. (1994). Late Eocene to Early Oligocene dinoflagellate cysts from the Priabonian type-area (northwest Italy): biostratigraphy and palaeoenvironmental interpretation. Palaeogeography, Palaeoclimatology, Palaeoecology, 107, 121–163.CrossRefGoogle Scholar
Brinkhuis, H. & Visscher, H. (1995). The upper boundary of the Eocene Series: a reappraisal based on dinoflagellate cyst biostratigraphy and sequence stratigraphy. In Geochronology, Time Scales and Global Stratigraphic Correlation, ed. Berggren, W. A., Kent, D. V., Aubry, M.-P. & Hardenbol, J.. Tulsa, SEPM Special Publication54, pp. 295–304.CrossRefGoogle Scholar
Brönnimann, P., and Resig, J. (1971). A Neogene Globigerinacean biochronologic timescale of the southwestern Pacific. Initial Reports of the Deep Sea Drilling Project, 7, 1235–1469.Google Scholar
Brotzen, F. (1959). On Tylocidaris species (Echinoidea) and the stratigraphy of the Danian of Sweden, with a bibliograp[hy of the Danian and the Paleocene. Sveriges Geologiska Undersökning, Series C, 54, 1–81.Google Scholar
Brown, C. W. & Stephenson, S. (1991). Geology of the Murray Basin. Canberra: Bureau of Mineral Resources, Bulletin 235.Google Scholar
Buckland, W. (1837). Geology and Mineralogy Considered with Reference to Natural Theology. London: William Pickering, 2 volumes.CrossRefGoogle Scholar
Buckman, S. S. (1902). The term ‘Hemera’. Geological Magazine, n.s., 9, 554–557.CrossRefGoogle Scholar
Buckman, S. S.(1903). The term ‘Hemera’. Geological Magazine, n.s., 10, 95–96.Google Scholar
Buffetaut, E. (1987). A Short History of Palaeontology. London: Wolfboro.Google Scholar
Butler, R. F. & Lindsay, E. H. (1985). Mineralogy of magnetic minerals and revised magnetic polarity stratigraphy of continental sediments, San Juan Basin, New Mexico. Journal of Geology, 93, 535–554.CrossRefGoogle Scholar
Butterfield, H. (1949). The Origins of Modern Science: 1300–1800. London: Bell, 217 pp.Google Scholar
Butterfield, H.(1951). The Whig Interpretation of History. London: G. Bell, 132 pp.Google Scholar
Buzas, M. A. & Culver, S. J. (1984). Species duration and evolution: Benthic foraminifera on the Atlantic continental margin of North America. Science, 225, 829–830.CrossRefGoogle ScholarPubMed
Buzas, M. A. & Culver, S. J.(1989). Biogeographic and evolutionary patterns of continental margin foraminifera: Paleobiology, 15(1), 11–19.CrossRefGoogle Scholar
Buzas, M. A. & Culver, S. J.(1998). Assembly, disassembly, and balance in marine paleocommunities. Palaios, 13, 263–275.CrossRefGoogle Scholar
Cain, A. J. (1954). Animal Species and their Evolution. London: Hutchinson's University Library, 190 pp.Google Scholar
Cain, A. J.(1960). Animal Species and their Evolution. New York: Harper & Brothers.Google Scholar
Callomon, J. H. (1995). Time from fossils: S. S. Buckman and Jurassic high-resolution geochronology. In Milestones in Geology, ed. Bas, M. J., Geological Society, London, Memoir, 16, 127–150.Google Scholar
Cande, S. C. & Kent, D. V. (1992). A new geomagnetic polarity time scale for the Late Cretaceous and Cenozoic. Journal of Geophysical Research, 97, 13917–13951.CrossRefGoogle Scholar
Cande, S. C. & Kent, D. V.(1995). Revised calibration of the geomagnetic polarity timescale for the Late Cretaceous and Cenozoic. Journal of Geophysical Research, 100, 6093–6095.CrossRefGoogle Scholar
Carney, J. L. & Pierce, R. W. (1995). Graphic correlation and composite standard databases as tools for the exploration biostratigrapher. In Graphic Correlation and the Composite Standard Approach, ed. Mann, K. O., Lane, H. R., & Stein, J. R., pp. 23–44. SEPM (Society for Sedimentary Geology), Special Publication No. 53.CrossRefGoogle Scholar
Caron, M. & Homewood, P. (1983). Evolution of early planktic foraminifers. Marine Micropaleontology, 7, 453–462.CrossRefGoogle Scholar
Carpenter, W. B., Parker, W. K. & Jones, T. R. (1862). Introduction to the Study of the Foraminifera. London: The Ray Society.Google Scholar
Carr, E. H. (1961). What is History? New York, N.Y: Random House, 209 pp.Google Scholar
Carter, A. N. (1958a). Pelagic foraminifera in the Tertiary of Victoria: Geological Magazine, 95, 297–304.CrossRefGoogle Scholar
Carter, A. N.(1958b). Tertiary foraminifera from the Aire District, Victoria: Geological Survey of Victoria Bulletin 55, 1–76.Google Scholar
Carter, R. M. (1974). A New Zealand case-study of the need for local time-scales. Lethaia, 7, 181–202.CrossRefGoogle Scholar
Carter, R. M. and Naish, T. R. (1998). Have local stages outlived their usefulness for the New Zealand Pliocene–Pleistocene? New Zealand Journal of Geology and Geophysics, 41, 271–279.CrossRefGoogle Scholar
Castradori, D. (2002). A complete standard chronostratigraphic scale: how to turn a dream into reality? Episodes, 25, 107–110.Google Scholar
Cavelier, C., Chateauneuf, J. J., Pomerol, C., Rabussier, D., Renard, M. & Vergnaud-Grassini, C. (1981). The geological events at the Eocene/Oligocene boundary. Palaeogeography, Palaeoclimatology, Palaeoecology, 36, 223–248.CrossRefGoogle Scholar
Chaisson, W. P. (2003). Vicarious living: Pliocene menardellids between an isthmus and an ice sheet. Geology, 31, 1085–8.CrossRefGoogle Scholar
Chamberlin, T. C. (1898). The ulterior basis of time divisions and the classification of geologic history. Journal of Geology, 6, 449–462.CrossRefGoogle Scholar
Chapman, M. R. (2000). The response of planktonic foraminifera to the Late Pliocene intensification of Northern Hemisphere glaciation. In Biotic Response to Global Change: The Past 145 Million Years, ed., Culver, S. J. & Rawson, P. F., pp. 79–96. London: British Museum of Natural History & Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Chaproniere, G. C. H. (1975). Palaeoecology of Oligo-Miocene Larger Foraminiferida, Australia. Alcheringa, 1, 37–58.CrossRefGoogle Scholar
Chaproniere, G. C. H.(1981). Australasian mid-Tertiary correlations, larger foraminiferal associations and their bearing on the East Indian Letter Classification. BMR Journal of Australian Geology and Geophysics, 6, 145–151. Canberra: Bureau of Mineral Resources (now Geoscience Australia).Google Scholar
Chaproniere, G. C. H.(1984). Oligocene and Miocene larger Foraminiferida from Australia and New Zealand. BMR Bulletin Australian Geology & Geophysics, 188, 1–98.Google Scholar
Chaproniere, G. C. H., Shafik, S., Truswell, E. M., Macphail, M. K. & Partridge, A. D. (1996). Ch. 2.10, Cainozoic (Chart 10). In An Australian Phanerozoic Timescale, ed. Young, G. C. & Laurie, J. R., pp. 175–186. Oxford: Oxford University Press.Google Scholar
Cifelli, R. (1969). Radiation of Cenozoic planktonic foraminifera. Systematic Zoology, 18, 154–168.CrossRefGoogle Scholar
Cifelli, R.(1990). A history of the classification of the foraminifera (1826–1933). Part I, Foraminiferal classification from d'Orbigny to Galloway. Cushman Foundation for Foraminiferal Research, Special Publication, 27, 1–88.Google Scholar
Cita, M. B. (1975). The Miocene–Pliocene boundary: history and definition. In Late Neogene Epoch Boundaries, ed. Saito, T & Burckle, L. H.. New York: Micropaleontology Press, Special Publication 1, 1–30.Google Scholar
Claridge, M. F., Dawah, H. A. & Wilson, M. R., Editors (1997). Species: The Units of Biodiversity. London: Chapman and Hall.Google Scholar
Clarke, W. J. & Blow, W. H. (1969). The inter-relationships of some late Eocene, Oligocene and Miocene larger foraminifera and plankton biostratigraphic indices. In Proceedings of the International Conference on Planktonic Microfossils (Geneva 1967), ed. Brönnimann, P. & Renz, H. H.. Leiden: E. J. Brill, vol. 2, 82–97.Google Scholar
Colin, J.-P. & Lethiers, F. (1988). The importance of ostracods in biostratigraphic analysis. In Ostracoda in the Earth Sciences, ed. Deckker, P., Colin, J.-P. & Peypouquet, J.-P., pp. 27–46. Amsterdam: Elsevier.Google Scholar
Conkin, B. M. & Conkin, J. E. (1984). Stratigraphy: Foundations and Concepts. Van Nostrand and Reinhold, New York, 335 pp.Google Scholar
Cooper, R. A., ed. (2004). The New Zealand geological timescale. Institute of Geological and Nuclear Sciences, Monograph 22.
Cooper, R. A., Crampton, J. S., Raine, J. I.et al. (2001). Quantitative biostratigraphy of the Taranaki Basin, New Zealand: a deterministic and probabilistic approach. American Association of Petroleum Geologists Bulletin, 85, 1469–1498.Google Scholar
Cooper, R. A., Crampton, J. S. & Uruski, C. I. (2000). The Time-Calibrated Composite: A Powerful Tool in Basin Exploration. Proceedings of the 2000 New Zealand Petroleum Conference, Christchurch, New Zealand, pp. 346–364. Wellington: Ministry of Commerce.Google Scholar
Corfield, R. M. & Cartlidge, J. E. (1991). Oceanographic and climatic implications of the Palaeocene carbon isotope maximum. Terra Nova, 4, 443–455.CrossRefGoogle Scholar
Corfield, R. M. & Norris, R. D. (1998). The oxygen and carbon isotopic context of the Paleocene/Eocene Epoch boundary. In Late Paleocene–early Eocene Climatic and Biotic Events in the Marine and Terrestrial Records, ed. Aubry, M.-P., Lucas, S. G. & Berggren, W. A., 124–137. New York: Columbia University Press.Google Scholar
Cowie, J. W. (1986). Guidelines for boundary stratotypes. Episodes, 9, 78–82.Google Scholar
Cowie, J. W.(1990). Global boundary stratotypes: overview. In Palaeobiology: A Synthesis, ed. Briggs, D. E. G. & Crowther, P. R., pp. 471–475. Oxford: Blackwell.Google Scholar
Cowie, J. W., Ziegler, W. & Remane, J. (1989). Stratigraphic commission accelerates progress. Episodes, 12, 79–83.Google Scholar
Cox, A., Doell, R. R. & Dalrymple, G. B. (1964). Geomagnetic polarity epochs. Science, 143: 351–352.CrossRefGoogle ScholarPubMed
Cracraft, J. (1987). Species concepts and the ontology of evolution. Biology and Philosophy, 2, 329–346.CrossRefGoogle Scholar
Crespin, I. (1943). The Stratigraphy of the Tertiary Marine Rocks in Gippsland, Victoria: Palaeontological Bulletin, 101 p. (mimeo) (Department of Supply and Shipping, Australia).Google Scholar
Croneis, C. (1941). Micropaleontology, past and future. American Association of Petroleum Geologists, Bulletin, 35, 1308–1355.Google Scholar
Cross, T. A. & Lessenger, M. A. (1988). Seismic stratigraphy. Annual Reviews of Earth and Planetary Sciences, 16, 319–354.CrossRefGoogle Scholar
Culver, S. J. & Buzas, M. A. (2000). Response of shallow water foraminiferal palaeocommunities to global and regional environmental change. In Biotic Response to Global Change: The Past 145 Million Years, ed. Culver, S. J., & Rawson, P. F., pp. 122–134. London: British Museum of Natural History & Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Cushman, J. A. (1927). Some new genera of the foraminifera. Contributions Cushman Laboratory Foraminiferal Research, 2, 77–81.Google Scholar
Cushman, J. A. & Stainforth, R. M. (1945). The foraminifera of the Cipero Marl Formation of Trinidad, British West Indies. Cushman Laboratory Foraminiferal Research Special Publication, 14, 1–74.Google Scholar
D'Hondt, S. (1991). Phylogenetic and stratigraphic analysis of earliest Paleocene triserial and biserial planktonic foraminifera. Journal of Foraminiferal Research, 21, 168–181.CrossRefGoogle Scholar
D'Hondt, S. & Zachos, J. C. (1993). On stable isotopic variation and earliest Paleocene planktic foraminifera. Paleoceanography, 8, 527–547.CrossRefGoogle Scholar
D'Hondt, S. & Zachos, J. C.(1998). Cretaceous foraminifera and the evolutionary history of planktic photosymbiosis. Paleobiology, 24, 512–523.CrossRefGoogle Scholar
Daniel, G. (1962). The Idea of Prehistory. London: Watts, 171 pp.Google Scholar
Darling, K. F., Kucera, M., Pudsey, C. J.et al. (2004). Molecular evidence links gyptic diversification in palar planktonic pratists to Quaternary chriate dynamics. Proceedings National Academy of Sciences, USA, 101, 7657–66.CrossRefGoogle Scholar
Darling, K. F., Wade, C. M., Kroon, D. & Leigh Brown, A. J. (1997). Planktonic foraminiferal molecular evolution and their polyphyletic origins from benthic taxa. Marine Micropaleontology, 30, 251–266.CrossRefGoogle Scholar
Darling, K. F., Wade, C. M., Kroon, D., Leigh Brown, A. J. & Bijma, J. (1999). The diversity and distribution of modern planktonic foraminiferal small subunit ribosomal RNA genotypes and their potential as tracers of past and present ocean circulations. Paleoceanography, 14, 3–12.CrossRefGoogle Scholar
Darling, K. F., Wade, C. M., Stewart, I. A.et al. (2000). Molecular evidence for genetic mixing of Arctic and Antarctic subpolar populations of planktonic foraminifers. Nature, 405, 43–47.CrossRefGoogle ScholarPubMed
Darragh, T. A. (1985). Molluscan biogeography and biostratigraphy of the Tertiary of southern Australia. Alcheringa 9, 83–116.CrossRefGoogle Scholar
Darwin, C. (1859, [1964]). On the Origin of Species: A Facsimile of the 1st Edition with an Introduction by Ernst Mayr. Cambridge, MA: Harvard University Press.Google Scholar
Davies, A. M. (1934). Tertiary Faunas: A Text-Book for Oilfield Palaeontologists and Students of Geology. London: Thomas Murby, 1934–35 [vol. 1, 1935], 2 volumes.Google Scholar
Vargas, C., Bonzon, M., Rees, N., Pawlowski, J., Zaninetti, L. (2002). A molecular approach to biodiversity and biogeography in the planktonic foraminifer Globigerinella siphonifera (d'Orbigny). Marine Micropaleontology, 45, 101–116.CrossRefGoogle Scholar
Vargas, C., Norris, R. D., Zaninetti, L., Gibb, S. W. & Pawlowski, J. (1999). Molecular evidence of cryptic speciation in planktonic foraminifers and their relation to oceanic provinces. Proceeding of the National Academy of Sciences USA, 96, 2864–2868.CrossRefGoogle ScholarPubMed
Vargas, C., Renaud, S., Hillbrecht, H. & Pawlowski, J. (2001). Pleistocene adaptive radiation in Eloborotalia transcatulinoides: genetic, morphologic, and environmental evidence. Paleobiology, 27, 104–2s.2.0.CO;2>CrossRefGoogle Scholar
Desmond, A. J. (1989). The Politics of Evolution: Morphology, Medicine, and Reform in Radical London. Chicago: University of Chicago Press, 503 pp.Google Scholar
Diener, C. (1925). Grundzüge der Biostratigraphie. Leipzig: Deuticke.Google Scholar
DiMichele, W. A. (1994). Ecological patterns in time and space. Paleobiology, 20, 89–92.CrossRefGoogle Scholar
DiMichele, W. A., Behrensmeyer, A. K., Olszewski, T. D.et al. (2004). Long-term stasis in ecological assemblages: evidence from the fossil record. Annual Review of Ecology and Systematics, 35, 285–322.CrossRefGoogle Scholar
Dobzhansky, T. (1937). Genetics and the Origin of Species. New York: Columbia University Press.Google Scholar
Donovan, D. T. (1966). Stratigraphy: An Introduction to Principles. London: Murby.Google Scholar
Dorf, E. (1955). Plants and the geological time scale. In Crust of the Earth, ed. Poldervaart, A.. Geological Society of America, Special Paper, 62, 575–592.Google Scholar
Dott, R. H. Jr, ed. (1992). Eustasy: the historical ups and downs of a major historical concept. Geological Society of America, Memoir, 180.Google Scholar
Douglas, R. G. & Savin, S. M. (1978). Oxygen isotope evidence for the depth stratification of Tertiary and Cretaceous planktonic foraminifera. Marine Micropaleontology 3, 175–196.CrossRefGoogle Scholar
Dowsett, H. J. (1988). Diachrony of late Neogene microfossils in the southwest Pacific Ocean: application of the graphic correlation method. Paleoceanography, 3, 209–222.CrossRefGoogle Scholar
Doyle, P. & Bennett, M. R. (Editors) (1997). Unlocking the Stratigraphical Record: Advances in Modern Stratigraphy. New York: John Wiley & Sons.Google Scholar
Drooger, C. W. (1956). Transatlantic correltion of the Oligo-Miocene by means of foraminifera. Micropaleontology, 2, 183–192.CrossRefGoogle Scholar
Drooger, C. W.(1963). Evolutionary trends in the Miogypsinidae. In Evolutionary Trends in Foraminifera, ed. Koenigswald, G. H. R., Emeis, J. D., Buning, W. L. & Wagner, C. W., pp. 315–349. Amsterdam: Elsevier Publishing Company.Google Scholar
Drooger, C. W.(1966). Zonation of the Miocene by means of planktonic foraminifera. With additional comments by Z. Reiss, M. B. Cita, W. H. Blow, F. E. Eames, R. M. Stainforth & H. M. Bolli. In Committee on Mediterranean Stratigraphy, ed. C. W. Drooger, Z. Reiss, R. F. Rutsch & P. Marks. Proceedings of the third session in Berne, 8–13 June 1964. International Union of Geological Sciences Commission on Stratigraphy, pp. 40–50, Leiden: E. J. Brill.
Drooger, C. W.(1974). The boundaries and limits of stratigraphy. Koninklijke Nederl. Akad. Wetenschappen, Amsterdam, Proc. Series B, 77, 159–176.Google Scholar
Drooger, C. W.(1993). Radial Foraminifera: Morphometrics and Evolution. Amsterdam, New York: North-Holland, 242 pp.Google Scholar
Durham, J. W. (1950). Cenozoic marine climates of the Pacific coast. Geological Society of America, Bulletin, 61, 243–1264.Google Scholar
Eames, F. E. & Savage, R. J. G. (1975). Tertiary Faunas: A Text-Book for Oilfield Palaeontologists and Students of Geology, by Davies, A. Morley, 2nd edition revised and brought up to date. London: Allen & Unwin, vols. 1, 2.Google Scholar
Eames, F. E., Banner, F. T., Blow, W. H. & Clarke, W. J. (1962). Fundamentals of Mid-Tertiary Stratigraphical Correlation. Cambridge: Cambridge University Press.Google Scholar
Edwards, L. E. (1982a). Quantitative biostratigraphy: the methods should suit the data. In Quantitative Stratigraphic Correlation, ed. Cubitt, J. M. & Reyment, R. A., pp. 45–60. New York: John Wiley.Google Scholar
Edwards, L. E.(1982b). Numerical and semi-objective biostratigraphy: review and predictions. Proceedings of the Third North American Palaeontological Convention, 1, 147–152.Google Scholar
Edwards, L. E.(1984). Insights on why the graphic correlation (Shaws method) works. Journal of Geology, 92, 583–597.CrossRefGoogle Scholar
Edwards, L. E.(1989). Supplemented graphic correlation: A powerful tool for paleontologists and nonpaleontologists. Palaios, 4, 127–143.CrossRefGoogle Scholar
Eicher, D. L. (1976). Geologic Time (2nd edn). Englewood Cliffs: Prentice-Hall.Google Scholar
Eldredge, N. (1979). Alternative approaches to evolutionary theory. Carnegie Museum Natural History Bulletin, 13, 7–19.Google Scholar
Eldredge, N.(1989). Macroevolutionary Dynamics: Species, Niches and Adaptive Peaks. New York: McGraw-Hill.Google Scholar
Eldredge, N. & Gould, S. J. (1972). Punctuated equilibria: an alternative to phyletic gradualism. In Models in Paleobiology, ed. Schopf, T. J. M., pp. 82–115. San Francisco: Freeman, Cooper.Google Scholar
Eldredge, N. & Gould, S. J.(1977). Evolutionary models and biostratigraphic strategies. In Concepts and Methods of Biostratigraphy, ed. Kauffman, E. G. & Hazel, J. E., pp. 3–22. Stroudsberg, PA: Dowden, Hutchison and Ross.Google Scholar
Emiliani, C. (1955). Pleistocene temperatures. Journal of Geology, 63, 538–578.CrossRefGoogle Scholar
Emiliani, C.(1969). A new paleontology. Micropaleontology, 15, 265–300.CrossRefGoogle Scholar
Emiliani, C.(1982). Extinctive evolution: extinctive and competitive evolution combine into a unified model of evolution. Journal of Theoretical Biology, 97, 13–33.CrossRefGoogle Scholar
Endler, J. (1989). In Speciation and its Consequences, ed. Otte, D. & Endler, J.. Sunderland, MA: Sinauer & Associates.Google Scholar
Ereshefsky, M., (ed.) (1992). The Units of Evolution: Essays on the Nature of Species. Cambridge, MA: The MIT Press.Google Scholar
Ereshefsky, M.(2001). The Poverty of the Linnaean Hierarchy: A Philosophical Study of Biological Taxonomy. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Ericson, D. B. (1961). Pleistocene climatic record in some deep sea sediment cores. Science, 95, 537–541.Google Scholar
Ericson, D. B. & Wollin, G. (1956). Micropaleontological and isotopic determinations of Pleistocene climates. Micropaleontology, 2, 257–270.CrossRefGoogle Scholar
Ericson, D. B. & Wollin, G.(1968). Pleistocene climates and chronology in deep sea sediments. Science, 162, 1227–1234.CrossRefGoogle ScholarPubMed
Evernden, J. F., Savage, D. E., Curtis, G. H. & James, G. T. (1964). Potassium-argon dates and the Cenozoic mammalian chronology of North America. American Journal of Science, 62, 145–198.CrossRefGoogle Scholar
Fejfar, O. & Heinrich, W.-D. (1989). Muroid rodent biochronology of the Neogene and Quaternary. In European Neogene Mammal Chronology, ed. Lindsay, E. H., Fahlbusch, V. & Mein, P., pp. 91–118. NATO Advanced Research Workshop on European Neogene Mammal chronology, Schloss Reisensburg. New York & London: Plenum Press.CrossRefGoogle Scholar
Finlay, H. J. (1947). The foraminiferal evidence for Tertiary trans-Tasman correlation. Transactions Royal Society New Zealand, 76, 327–352.Google Scholar
Finlay, H. J. & Marwick, J. A. (1940). The divisions of the upper Cretaceous and Tertiary in New Zealand. Transactions of the Royal Society of New Zealand, 70, 77–135.Google Scholar
Finlay, H. J. & Marwick, J. A.(1947). New divisions of the New Zealand upper Cretaceous and Tertiary. New Zealand Journal of Science & Technology, B28, 228–236.Google Scholar
Fornaciari, E. & Rio, D. (1996). Latest Oligocene to early Miocene quantitative calcareous nannofossil biostratigraphy in the Mediterranean region. Micropaleontology, 42, 1–26.CrossRefGoogle Scholar
Fuller, S. (2003). Kuhn vs Popper: The Struggle for the Soul of Science. Icon Books, UK.Google Scholar
Fischer, A. G. (1980). Gilbert: bedding rhythms and geochronology. In The Scientific Ideas of G. K. Gilbert, ed. Yochelson, E., pp. 93–104. Geological Society of America, Special Paper 183.Google Scholar
Fischer, A. G.(1981). Climatic oscillations in the biosphere. In Biotic Crises in Ecological and Evolutionary Time, ed. Nitecki, M. H., pp. 103–121. New York: Academic Press.Google Scholar
Fischer, A. G.(1984). The two Phanerozoic supercycles. In Catastrophes and Earth History, ed. Berggren, W. A. & Couvering, J. A., pp. 9–34. Princeton: Princeton University Press.CrossRefGoogle Scholar
Fischer, A. G.(1986). Climate rhythms recorded in strata. Annual Review of Earth & Planetary Sciences, 14, 351–376.CrossRefGoogle Scholar
Fischer, A. G. & Arthur, M. A. (1977). Secular variations in the pelagic realm. In Deep-Water Carbonate Environments, ed. Cook, H. & Enos, P.. Society of Economic Paleontologists and Mineralogists, Special Publication, 25, 19–25.CrossRefGoogle Scholar
Fischer, A. G. & Herbert, T. (1986). Stratification rhythms: Italo-American studies in the Umbrian facies. Mem. Geol. Soc. Italia, 31, 45–51.Google Scholar
Fischer, A. G., Herbert, T. D., Napoleone, G., Premoli Silva, I. & Ripepe, R. (1991). Albian pelagic rhythms (Piobbico core). Journal of Sedimentary Petrology, 62, 1146–1172.Google Scholar
Fisher, D. C. (1994). Stratocladistics: morphological and temporal patterns and their relation to phylogenetic process. In Interpreting the Hierarchy of Nature: From Systematic Patterns to Evolutionary Process Theories, ed. Grande, L. & Rieppel, O.. pp. 133–171, Orlando: Academic Press.Google Scholar
Flower, B. P. & Kennett, J. P. (1993) Relations between Monterey Formation deposition and global cooling. Geology, 88, 10–11.Google Scholar
Flower, B. P. & Kennett, J. P.(1995) Middle Miocene deepwater paleoceanography in the southwest Pacific: relations with East Antarctic ice sheet development. Paleoceanography, 10, 1095–1113.CrossRefGoogle Scholar
Fordham, B. G. (1986). Miocene–Pleistocene foraminifers from DSDP Sites 208 and 77, and phylogeny and classification of Cenozoic species. Evolutionary Monographs, 6, 200 pp.Google Scholar
Fortey, R. A. (1993). Charles Lapworth and the biostratigraphic paradigm. In Milestones in Geology, ed. Bas, M. J., Geological Society, London, Memoir, 16, 93–104.Google Scholar
Frerichs, W. E. (1971). Evolution of planktonic foraminifera and paleotemperatures. Journal of Paleontology, 45, 963–968.Google Scholar
Funnell, B. M. (1964). The Tertiary Period. In The Phanerozoic time scale: a symposium. Quarterly Journal Geological Society London, 120(S), 179–191. [no editor named]Google Scholar
Gallagher, S. J. & Holdgate, G. R. (1996). Sequence Stratigraphy and Biostratigraphy of the Onshore Gippsland Basin, S. E. Australia. Australian Sedimentologists Group Field Guide Series No. 11, Geological Society of Australia, 70 pp.
Gardner, J. (1931). Relation of certain foreign faunas to Midway fauna of Texas. American Association of Petroleum Geologists, Bulletin, 15, 149–160.Google Scholar
Gayon, J. (1990). Critics and criticisms of the modern synthesis: the viewpoint of a philosopher. Evolutionary Biology, 24, 1–49.Google Scholar
Geikie, A. (1905). The Founders of Geology. London: Macmillan.Google Scholar
George, T. N. (1956). Biospecies, chronospecies and morphospecies. Systematics Association Publication No. 2, 123–137.Google Scholar
George, T. N.et al. (1967). The stratigraphical code – Report of the Stratigraphical Code Subcommittee. Geological Society of London, Proceedings, 1638, 75–87.Google Scholar
George, T. N.et al. (1969). Recommendations on stratigraphical usage. Geological Society of London, Proceedings, 1638, 139–166 (2nd revision of 1967 Report of the Stratigraphical Code Subcommittee).Google Scholar
Ghiselin, M. T. (1974). A radical solution to the species problem. Systematic Zoology, 23, 536–544.CrossRefGoogle Scholar
Ghiselin, M. T.(1997). Metaphysics and the Origin of Species. Albany: State University of New York Press, 377 pp.Google Scholar
Gignoux, M. (1955). Stratigraphic Geology (English translation by G. G. Woodford of the 1950 French edition of ‘Géologie Stratigraphique’). San Francisco: W. H. Freeman, 682 pp.Google Scholar
Gili, E., Skelton, P. W., Vicens, E. & Obrador, A. (1995). Corals to rudists – an environmentally induced assemblage succession. Palaeogeography, Palaeoclimatology, Palaeoecology, 119, 127–136.CrossRefGoogle Scholar
Gillispie, C. C. (1951). Genesis and Geology: A Study in the Relations of Scientific Thought, Natural Theology, and Social Opinion in Great Britain, 1790–1850. Cambridge, MA: Harvard University Press.Google Scholar
Gilluly, J. (1949). Distribution of mountain building in geologic time. Geological Society of America, Bulletin, 60, 561–590.CrossRefGoogle Scholar
Gingerich, P. D. (1979). Stratophenetic approach to phylogeny reconstruction in vertebrate paleontology. In Phylogenetic Analysis and Paleontology, ed. Cracraft, J. & Eldredge, N., pp. 41–79. New York: Columbia University Press.Google Scholar
Gingerich, P. D.(1990). Stratophenetics. In Paleobiology: A Synthesis, ed. Briggs, D. E. G. & Crowther, P. R., pp. 437–441. Oxford: Blackwells Scientific Publishing.Google Scholar
Glaessner, M. F. (1937). Planktonforaminiferen aus der Kreide und dem Eozän und ihre stratigraphische Bedeutung. Studies in Micropaleontology, 1, 27–52.Google Scholar
Glaessner, M. F.(1943). Problems of stratigraphic correlation in the Indo-Pacific region. Proceedings of the Royal Society of Victoria, 55, 41–80.Google Scholar
Glaessner, M. F.(1945). Principles of Micropalaeontology. Melbourne: Melbourne University Press, 296 pp.Google Scholar
Glaessner, M. F.(1951). Three foraminiferal zones in the Tertiary of Australia. Geological Magazine, 88, 273–283.CrossRefGoogle Scholar
Glaessner, M. F.(1953). Time-stratigraphy and the Miocene Epoch. Geological Society of America, Bulletin, 64, 647–658.CrossRefGoogle Scholar
Glaessner, M. F.(1955). Taxonomic, stratigraphic and ecologic studies of foraminifera and their interrelations. Micropaleontology, 1, 3–8.CrossRefGoogle Scholar
Glaessner, M. F.(1966). Problems of palaeontology. Journal of the Geological Society of India, 7, 14–27.Google Scholar
Glaessner, M. F.(1967). Time scales and Tertiary correlations. In Tertiary Correlations and Climatic Changes in the Pacific, ed. Hatai, K., 11th Pacific Science Congress, Tokyo, 1966, Symposium No. 25, 1–5.Google Scholar
Glaessner, M. F. & Wade, M. (1958). The St. Vincent Basin. In Geology of South Australia, ed. Glaessner, M. F. & Parkin, L. W.. Journal of the Geological Society of Australia, 5(2), 115–126.Google Scholar
Goldstein, S. (1999). Foraminifera: a biological overview. In Modern Foraminifera, ed. Gupta, B. K. Sen, pp. 37–56. Dordrecht: Kluwer Academic Publishers.Google Scholar
Gould, S. J. (1980). Is a new and general theory of evolution emerging? Paleobiology, 6, 119–130.CrossRefGoogle Scholar
Gould, S. J.(1982). Darwinism and the expansion of evolutionary theory. Science, 216, 380–387.CrossRefGoogle ScholarPubMed
Gould, S. J.(1984). Toward the vindication of punctuational change. In Catastrophes and Earth History, ed. Berggren, W. A. & Couvering, J. A., pp. 9–34. Princeton: Princeton University Press.CrossRefGoogle Scholar
Gould, S. J.(1985). The paradox of the first tier: an agenda for paleobiology. Paleobiology, 11, 2–12.CrossRefGoogle Scholar
Gould, S. J.(1986). Evolution and the triumph of homology, or why history matters. American Scientist, Jan–Feb, 60–69.Google Scholar
Gould, S. J.(1987). Time's Arrow, Time Cycle. Cambridge, MA: Harvard University Press.Google Scholar
Gould, S. J.(1996). Life's Grandeur: The Spread of Excellence from Plato to Darwin. London: Jonathan Cape.CrossRefGoogle Scholar
Gould, S. J.(2002). The Structure of Evolutionary Theory. Cambridge, MA, & London: Belknap Press of Harvard University Press.Google Scholar
Gould, S. J. & Eldredge, N. (1977). Punctuated equilibria; the tempo and mode of evolution reconsidered. Paleobiology, 3, 115–151.CrossRefGoogle Scholar
Gould, S. J. & Eldredge, N.(1993). Punctuated equilibrium comes of age. Nature, 366, 223–227.CrossRefGoogle ScholarPubMed
Gould, S. J. & Lewontin, R. W. (1979). The spandrels of San Marco and the Panglossian paradigm: a critique of the adaptationist programme. Proceedings of the Royal Society of London, Series B, 205, 581–598.CrossRefGoogle Scholar
Grabau, A. W. (1940). The Rhythm of the Ages. Peking: Henri Vetch Publishers.Google Scholar
Gradstein, F. M. & Agterberg, F. P. (1982). Models of Cenozoic foraminiferal stratigraphy – northwestern Atlantic margin. In Quantitative Stratigraphic Correlation, ed. Cubitt, J. M. & Reyment, R. A., pp. 119–173. New York: John Wiley.Google Scholar
Gradstein, F. M. & Agterberg, F. P.(1985). Quantitative correlation in exploration micropaleontology. In Quantitative Stratigraphy, ed. Gradstein, F. M, Agterberg, F. P., Brower, J. C. & Schwarzacher, W. S., pp. 309–357. Paris: UNESCO & D. Reidel Publishing Company.Google Scholar
Gregory, T. R. (2004). Macroevalution, hierarchy theory, and the C-value sigma. Paleobiology 30:179–202.2.0.CO;2>CrossRefGoogle Scholar
Grünbaum, A. (1963). Philosophical Problems of Space and Time. New York: Alfred A. Knopf, 448 pp.Google Scholar
Guex, J. (1991). Biochronological Correlations. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Hacohen, M. H. (2000). Karl Popper, the Formative Years 1902–1945: Politics and Philosophy in Interwar Vienna. Cambridge University Press.Google Scholar
Hailwood, E. A. (1989). The role of magnetostratigraphy in the development of geological time scales. Paleoceanography, 4, 1–18.CrossRefGoogle Scholar
Hallam, A. (1983). Plate tectonics and evolution. In Evolution from Molecules to Men, ed. Bendall, D. S., pp. 367–386. Cambridge: Cambridge University Press.Google Scholar
Hallam, A.(1989). Great Geological Controversies, Second edition. Oxford: Oxford University Press.Google Scholar
Hallam, A.(1993). Phanerozoic Sea Level Changes. New York: Columbia University Press.Google Scholar
Hallam, A.(1985). Why are larger foraminifera large? Paleobiology, 11, 195–208.Google Scholar
Hallam, A.(1987). Fluctuations in the trophic resource continuum: a factor in global diversity cycles? Paleoceanography 2, 457–471.Google Scholar
Hallam, A.(1999). Symbiont-bearing foraminifera. In Modern Foraminifera, ed. Gupta, B. K. Sen, pp. 123–139. Dordrecht: Kluwer Academic Publishers.Google Scholar
Hallock, P., Premoli Silva, I. & Boersma, A. (1991). Similarities between planktonic and larger foraminiferal evolutionary trends through Paleogene paleoceanographic changes. Palaeogeography, Palaeoclimatology, Palaeoecology, 83, 49–64.CrossRefGoogle Scholar
Hancock, J. M. (1977). The historic development of concepts of biostratigraphic correlation. In Concepts and Methods of Biostratigraphy, ed. Kauffman, E. G. & Hazel, J. E., pp. 3–22. Stroudsberg, PA: Dowden, Hutchison & Ross.Google Scholar
Hansen, T. A. (1988). Early Tertiary radiation of marine molluscs and the long-term effects of the Cretaceous–Tertiary extinction. Paleobiology, 14, 37–51.CrossRefGoogle Scholar
Haq, B. U. (1982). Climatic acme events in the sea and on land. In Climate and Earth History, pp. 126–132. Washington: National Academy Press.Google Scholar
Haq, B. U. & Boersma, A., Editors (1978). Introduction to Marine Micropaleontology. New York: Elsevier.Google Scholar
Haq, B. U., Hardenbol, J. & Vail, P. R. (1987). The chronology of fluctuating sea level since the Jurassic. Science, 235, 1156–67.CrossRefGoogle Scholar
Haq, B. U., Hardenbol, J. & Vail, P. R.(1988). Mesozoic and Cenozoic chronostratigraphy and cycles of sea-level change. In Sea-Level Changes, an Integrated Approach, ed. Wilgus, C. K., Hastings, B. S., Kendall, C. G. S. C., Posamentier, H. W., Ross, C. A., & Wagoner, J. C., Society of Economic Paleontologists and Mineralogists, Special Publication, 42, 71–108.CrossRefGoogle Scholar
Hardenbol, J. & Berggren, W. A. (1978). A new Paleogene numerical time scale. In Contributions to the Geological Time Scale, ed. Cohee, G. V., Glaessner, M. F. & Hedberg, H. D.. American Association of Petroleum Geologists, Studies in Geology, 6, 216–234.Google Scholar
Hardenbol, J., Thierry, J., Farley, M. B., Jacquin, T., de Graciansky, P.-C. & Vail, P. R. (1998). Mesozoic and Cenozoic sequence chronostratigraphic framework of European basins. In Mesozoic and Cenozoic Sequence Stratigraphy of European Basins, ed. Graciansky, P.-C., Hardenbol, J., Jacquin, T. & Vail, P. R., SEPM (Society of Sedimentary Geology) Special Publication No. 60, 3–13.CrossRefGoogle Scholar
Harland, W. B. (1973). Stratigraphic classification, terminology and usage – essay review of ‘Hedberg, H. D., Editor, 1972. An international guide to stratigraphic classification, terminology and usage, Introduction and summary’. Geological Magazine, 110, 567–574.CrossRefGoogle Scholar
Harland, W. B.(1975). The two geological time scales. Nature, 253, 505–507.CrossRefGoogle Scholar
Harland, W. B.(1992). Stratigraphic regulation and guidance: a critique of current tendencies in stratigraphic codes and guides. Geological Society of America Bulletin, 104, 1231–1235.2.3.CO;2>CrossRefGoogle Scholar
Harland, W. B., Armstrong, R. L., Cox, A. V., Craig, L. E., Smith, A. G. & Smith, D. G. (1990). A Geological Time Scale 1989. Cambridge: Cambridge University Press, 163 pp.Google Scholar
Hart, M. B. (1980). A water depth model for the evolution of the planktonic Foraminiferida. Nature, 286, 252–254.CrossRefGoogle Scholar
Hart, M. B.(1990). Major evolutionary radiations of the planktonic foraminiferida. In Major Evolutionary Radiations, ed. Taylor, P. D. & Larwood, G. P.. Systematics Association Special Volume No. 42, 59–72.Google Scholar
Harzhauser, M. & Piller, W. E. (2004). Integrated stratigraphy of the Sarmatian (upper Middle Miocene) in the western central Paratethys. Stratigraphy, 1, 65–86.Google Scholar
Hay, W. W. (1972). Probabilistic stratigraphy. Ecologae Geologicae Helvetiae, 65, 255–266.Google Scholar
Hay, W. W. & Mohler, H. P. (1969). Paleocene–Eocene calcareous nannoplankton and high-resolution biostratigraphy. In Proceedings of the First International Conference on Planktonic Microfossils, ed. Brönnimann, P. & Renz, H. H., pp. 250–253. Leiden: E. J. Brill.Google Scholar
Hay, W. W. & Southam, J. R. (1978). Quantifying biostratigraphic correlation. Annual Reviews of Earth and Planetary Sciences, 6, 353–375.CrossRefGoogle Scholar
Haynes, J. (1981). Foraminifera. London: Macmillan.CrossRefGoogle Scholar
Hays, J. D., Imbrie, J. & Shackleton, N. J. (1976). Variations in the earth's orbit: pacemaker of the ice ages. Sciences, 194, 1121–1132.CrossRefGoogle ScholarPubMed
Hazel, J. E. (1989). Chronostratigraphy of Upper Eocene microspherules. Palaios, 4, 318–329.CrossRefGoogle Scholar
Hazel, J. E.(1993). Biostratigraphy. In Fossil Prokaryotes and Protists, ed. Lipps, J. H., pp. 44–50. Boston: Blackwell Scientific Publishers.Google Scholar
Heath, R. S. & McGowran, B. (1984). Neogene datum planes: foraminiferal successions in Australia with reference sections from the Ninety-east Ridge and the Ontong–Java Plateau. In Pacific Neogene Datum Planes: Contributions to Biostratigraphy and Chronology, ed. Ikebe, N. & Tsuchi, R., pp. 187–192. Tokyo: University of Tokyo Press.Google Scholar
Hedberg, H. D. (1948). Time-stratigraphic classification of sedimentary rocks. Geological Society of America, Bulletin 59, 447–462.CrossRefGoogle Scholar
Hedberg, H. D.(1959). Toward harmony in stratigraphic classification. American Journal of Science, 257, 674–683.CrossRefGoogle Scholar
Hedberg, H. D.(1961). The stratigraphic panorama (an inquiry into the bases for age determination and age classification of the earth's rock strata). Geological Society of America, Bulletin 72, 499–518.Google Scholar
Hedberg, H. D.ed. (1976). International Stratigraphic Guide: A Guide to Stratigraphic Classification, Terminology and Procedure. New York: John Wiley & Sons, 200 pp.Google Scholar
Heezen, B. C. & MacGregor, I. (1973). The evolution of the Pacific. Scientific American, 229, 102–112.CrossRefGoogle Scholar
Heezen, B. C., MacGregor, I., Foreman, H. P., Forristall, G. Z., Hekel, H., Hesse, , Hoskins, R. H., Jones, E. J. W., Krasheninnikov, V., Okada, H. & Ruff, M. H. (1973). Diachronous deposits: a kinematic interpretation of the post-Jurassic sedimentary sequence on the Pacific plate. Nature, 241, 25–32.CrossRefGoogle Scholar
Heirtzler, J. R., Dickson, G. O., Herron, E. M., Pitman, W. C. & Pichon, X. (1968). Marine magnetic anomalies, geomagnetic field reversals, and motions of the ocean floor and continents. Journal Geophysical Research, 73, 2119–2136.CrossRefGoogle Scholar
Hemleben, Ch., Spindler, M. & Anderson, O. R. (1989). Modern Planktonic Foraminifera. New York: Springer-Verlag, 363 pp.CrossRefGoogle Scholar
Herbert, T. D. (1999). Toward a composite orbital chronology for the late Cretaceous and early Palaeocene GPTS. Philosophical Transactions Royal Society, 357, 1891–1905.CrossRefGoogle Scholar
Herbert, T. D. & D'Hondt, S. L. (1990). Precessional climate cyclicity in late Cretaceous–early Tertiary marine sediments: a high resolution chronometer of Cretaceous–Tertiary boundary events. Earth & Planetary Science Letters, 99, 263–275.CrossRefGoogle Scholar
Herbert, T. D. & Fischer, A. G. (1986). Milankovitch climatic origin of mid-Cretaceous black shale rhythms in central Italy. Nature, 321, 739–743.CrossRefGoogle Scholar
Herbert, T. D., Gee, J. S. & DiDonna, S. (1999). Precessional climatic cycles in the late Cretaceous South Atlantic: Long-term consequences of high-frequency variations. In Late Cretaceous Climates, ed. Barrera, E. & Johnson, C.. Boulder: Geological Society of America Special Volume 322, 105–120.Google Scholar
Herbert, T. D., Premoli Silva, I., Erba, E. & Fischer, A. G. (1995). Orbital chronology of Cretaceous–Paleocene marine sediments. In Geochronology Time Scales and Global Stratigraphic Correlation, ed. Berggren, W. A., Kent, D. V, Aubry, M.-P. & Hardenbol, J.. Tulsa: SEPM (Society of Sedimentary Geology) Special Publication 54, 81–94.CrossRefGoogle Scholar
Herbert, T. D., Stallard, R. F & Fischer, A. G. (1986). Anoxic events, productivity rhythms, and the orbital signature in a mid-Cretaceous pelagic core. Paleoceanography, 1, 495–506.CrossRefGoogle Scholar
Hess, J., Stott, L. D., Bender, M. L., Kennett, J. P. & Schilling, J. G. (1989). The Oligocene marine microfossil record: age assessments using strontium isotopes. Paleoceanography, 4, 655–679.CrossRefGoogle Scholar
Hesse, R., Foreman, H. P., Forristall, G. Z., Heezen, B. C., Hekel, H., Hoskins, R. H., Jones, E. J. W., Kaneps, A. G., Krasheninnikov, V., MacGregor, I. & Okada, H. (1974). Walther's facies rule in pelagic realm – a large-scale example from the Mesozoic–Cenozoic Pacific. Zeitschrift Deutsches Geologisches Gesellschaft, 125, 151–172.Google Scholar
Hickey, L. J., West, R. M., Dawson, M. R. & Choi, D. K. (1983). Arctic terrestrial biota: paleomagnetic evidence of age disparity with mid–northern latitudes during the Late Cretaceous and Early Tertiary. Science, 221, 153–1154.CrossRefGoogle ScholarPubMed
Hilgen, F. J. (1991). Astronomical calibration of Gauss to Matuyama sapropels in the Mediterranean and implications for the geomagnetic polarity time scale. Earth and Planetary Science Letters, 104, 226–244.CrossRefGoogle Scholar
Hilgen, F. J.(1994). An astronomically calibrated (polarity) time scale for the Pliocene–Pleistocene: a brief review. International Association of Sedimentologists, Special Publications, 19, 109–116.Google Scholar
Hilgen, F. J. & Langereis, C. G. (1993). A critical re-evaluation of the Miocene/Pliocene boundary as defined in the Mediterranean. Earth & Planetary Science Letters, 118, 167–179.CrossRefGoogle Scholar
Hilgen, F. J. & Langereis, C. G.(1994). Reply to comment on ‘A critical re-evaluation of the Miocene–Pliocene boundary as defined in the Mediterranean’, by R. H. Benson & D. A. Hodell. Earth and Planetary Science Letters, 118, 124, 251–254.CrossRefGoogle Scholar
Hilgen, F. J. & Krijgsman, W. (1999). Cyclostratigraphy and astrochronology of the Tripoli diatomite formation (pre-evaporite Messinian, Sicily, Italy). Terra Nova, 11, 16–22.CrossRefGoogle Scholar
Hilgen, F. J., Aziz, H. A., Krijgsman, W.et al. (1999). Present status of the astronomical (polarity) time-scale for the Mediterranean Neogene. Philosophical Transactions of the Royal Society of London A, 357, 1931–1947.CrossRefGoogle Scholar
Hilgen, F. J., Krijgsman, W., Langereis, C. G. & Lourens, L. J. (1997). Breakthrough made in dating of the geological record. Eos, 78, 285–292.CrossRefGoogle Scholar
Hilgen, F. J., Krijgsman, W., Raffi, I., Turco, E. & Zachariasse, (2000). Integrated stratigraphy and astronomical calibration of the Serravallian–Tortonian boundary section at Monte Gibliscemi (Sicily, Italy). Marine Micropaleontology, 38, 181–211.CrossRefGoogle Scholar
Hilgen, F. J., Lourens, L. J., Berger, A. & Loutre, M. F. (1993). Evaluation of the astronomically calibrated time scale for the late Pliocene and earliest Pleistocene. Paleoceanography, 8, 549–566.CrossRefGoogle Scholar
Hills, S. J. & Thierstein, H. R. (1989). Plio-Pleistocene calcareous plankton biochronology. Marine Micropaleontology, 14, 67–96.CrossRefGoogle Scholar
Hodell, D. A. (1994). Editorial: progress and paradox in strontium isotope stratigraphy. Paleoceanography, 9, 395.CrossRefGoogle Scholar
Hodell, D. A., Curtis, J. H., Sierro, F. J. & Raymo, M. E. (2001). Correlation of late Miocene to early Pliocene sequences between the Mediterranean and North Atlantic. Paleoceanography, 16, 164–178.CrossRefGoogle Scholar
Hodell, D. A. & Kennett, J. P. (1986). Late Miocene–early Pliocene stratigraphy and paleoceanography of the South Atlantic and southwest Pacific Oceans: a synthesis. Paleoceanography, 1, 285–311.CrossRefGoogle Scholar
Hodell, D. A. & Vayavananda, A. (1993). Middle Miocene paleoceanography of the western equatorial Pacific (DSDP site 289) and the evolution of Globorotalia (Fohsella). Marine Micropaleontology, 22, 279–310.CrossRefGoogle Scholar
Hodell, D. A. & Woodruff, F. (1994). Variations in the strontium isotopic ratio of sea water during the Miocene: stratigraphic and geochemical implications. Paleoceanography, 9, 405–426.CrossRefGoogle Scholar
Hodge, M. J. S. (1989). Darwin's theory and Darwin's argument. In What the Philosophy of Biology Is: Essays Dedicated to David Hull, ed. Ruse, M., pp. 163–182. Dordrecht: Kluwer Academic, 337 pp.CrossRefGoogle Scholar
Hoffman, A. (1983). Paleobiology at the crossroads: a critique of some modern paleobiological research programs. In Dimensions of Darwinism, ed. Grene, M., pp. 241–271. Cambridge: Cambridge University Press.Google Scholar
Hoffman, A.(1989). Arguments on Evolution: A Paleontologist's Perspective. New York, Oxford: Oxford University Press, 274 pp.Google Scholar
Hoffmann, A. & Reif, W.-E. (1988). On methodology of the biological sciences: from an evolutionary biological perspective. Neues Jahrbuch Geologie & Paläontologie, Abhandlungen, 177, 185–211.Google Scholar
Holdgate, G. R & Gallagher, S. J. (1997). Microfossil paleoenvironments and sequence stratigraphy of Tertiary cool-water carbonates, onshore Gippsland Basin, S. E. Australia. In Cool and Temperate Water Carbonates, ed. James, N. & Clarke, J. D. A.. Tulsa: Special Publication of the Society of Economic Paleontologists and Mineralogists, 56, 205–220.CrossRefGoogle Scholar
Holland, C. H. (1978). Stratigraphical classification and all that. Lethaia, 11, 85–90.CrossRefGoogle Scholar
Holland, C. H.(1990). Biostratigraphic units and the stratorype/golden spike concept. In Paleobiology: A Synthesis, ed. Briggs, D. E G. & Crowther, P. R., 461–465. Oxford: Blackwells Scientific Publishing.Google Scholar
Holland, S. M. (1995). Depositional sequences, facies control and the distribution of fossils. In Sequence Stratigraphy and Depositional Response to Eustatic, Tectonic and Climatic Forcing, ed. Haq, B. U.. pp. 1–23, Dordrecht: Kluwer Academic Publishers.CrossRefGoogle Scholar
Holland, S. M.(1999). The new stratigraphy and its promise for paleobiology. Paleobiology, 25, 409–416.CrossRefGoogle Scholar
Holland, S. M.(2000). The quality of the fossil record: a sequence stratigraphic perspective. In Deep time: Paleobiology's Perspective, ed. D. H. Erwin & S. L. Wing, Paleobiology, Supplement to vol. 26(4), 148–168.CrossRef
Hornibrook, N. deB. (1958). New Zealand Foraminifera: Key species in stratigraphy – No.6. New Zealand Journal Geology & Geophysics, 1, 653–676.CrossRefGoogle Scholar
Hornibrook, N. deB. (1965). A viewpoint on stages and zones. New Zealand Journal of Geology & Geophysics, 8, 1195–1212.CrossRefGoogle Scholar
Hornibrook, N. deB.(1969). Report on a Visit to the U.S.A. to Attend the I.U.G.S. Working Group for a Biostratigraphic Zonation of the Cretaceous and the Cenozoic. Lower Hutt: New Zealand Geological Survey Report 42, 1–11.
Hornibrook, N. deB. (1971). Inherent instability of biostratigraphic zonal schemes. New Zealand Journal of Geology & Geophysics, 14, 727–733.CrossRefGoogle Scholar
Hornibrook, N. deB. (1976). Jurassic, Cretaceous and Cenozoic stages divisions and zones used in New Zealand. United Nations Mineral Resources Development Series, 42, 81–93.Google Scholar
Hornibrook, N. deB. & Edwards, A. R. (1971). Integrated planktonic foraminiferal and calcareous nannoplankton datum levels in the New Zealand Cenozoic. In Proceedings of the II Planktonics Conference, Roma 1970, ed. Farinacci, A., 649–657. Roma: Edizioni Technoscienza.Google Scholar
Hornibrook, N. deB., Brazier, R. C. & Strong, C. P. (1989). Manual of New Zealand Permian to Pleistocene Foraminiferal Biostratigraphy. Lower Hutt, New Zealand Geological Survey Paleontological Bulletin 56, 175 pp.
Hottinger, L. (1981). The Resolution Power of the Biostratigraphic Clock Based on Evolution and its Limits. Barcelona: Universidad Barcelona, International Symposium on Concepts and Methods in Palaeontology, pp. 233–242.Google Scholar
Hottinger, L.(1982). Larger foraminifera, giant cells with a historical background. Naturwissenschaft, 69, 361–371.CrossRefGoogle Scholar
Hottinger, L.(1983). Processes determining the distribution of larger foraminifera in space and time. Utrecht Micropaleontological Bulletins, 30, 239–253.Google Scholar
Hottinger, L.(1990). Significance of diversity in shallow benthic foraminifera. Atti del Quarto Simposio Ecologia e Paleoecologia Communità Benthonice, Sorrento, 1–5 November 1988. Museo Regionale di Scienze Natural – Torino, pp. 35–51.
Hottinger, L.(1996). Sels nutritifs et biosédimentation. Mémoir de la Societé géologique de France, n.s., 169, 99–107.Google Scholar
Hottinger, L.(1997). Shallow benthic foraminiferal assemblages as signals for depth of their deposition and their limitations. Bulletin de la Societé géologique de France, 168, no. 4, 491–505.Google Scholar
Hottinger, L.(1998). Shallow benthic foraminifera at the Paleocene–Eocene boundary. Strata, 9, 61–64.Google Scholar
House, M. (1985). A new approach to an absolute timescale from measurements of orbital cycles and sedimentary microrhythms. Nature, 316, 721–725.CrossRefGoogle Scholar
Huber, B. T., Bijma, J. & Darling, K. (1997). Cryptic speciation in the living planktonic foraminifer Globigerinella siphonifera (d'Orbigny). Paleobiology, 23, 33–62.CrossRefGoogle Scholar
Hughes, N. F. (1989). Fossils as Information: New Recording and Stratal Correlation Techniques. Cambridge, New York: Cambridge University Press, 136 pp.CrossRefGoogle Scholar
Hulburt, R. C. Jr (1993). Taxic evolution in North American Neogene horses (Subfamily Equinae): the rise and fall of an adaptive radiation. Paleobiology, 19, 216–234.CrossRefGoogle Scholar
Hull, D. (1983). Popper and Plato's metaphor. In Advances in Cladistics, Vol. 2, ed. Platnick, N. & Funk, V. A., pp. 177–189. New York: Columbia University Press.Google Scholar
Hull, D.(1984). Historical entities and historical narratives. In Minds Machines and Evolution, ed. Hookway, C., pp. 17–42. Cambridge: Cambridge University Press.Google Scholar
Hull, D.(1988). Science as a Process: An Evolutionary Account of the Social and Conceptual Development of Science. Chicago: University of Chicago Press, 583 pp.CrossRefGoogle Scholar
Hull, D.(1989). A function for actual examples in philosophy of science. In What the Philosophy of Biology Is: Essays Dedicated to David Hull, ed. Ruse, M., pp. 309–322. Dordrecht: Kluwer Academic Publishers.CrossRefGoogle Scholar
Hull, D. L. (1997). The ideal species concept – and why we can't get it. In Species: The Units of Biodiversity, ed. Claridge, M. F., Dawah, H. A. & , M. R., pp. 357–380. Chapman & Hall.Google Scholar
Hunter, R. S. T., Arnold, A. J. & Parker, W. C. (1988). Evolution and homeomorphy in the development of the Paleocene Planorotalites pseudomenardii and the Miocene Globorotalia (Globorotalia) margaritae lineages. Micropaleontology, 31, 181–192.CrossRefGoogle Scholar
Huxley, J. S., ed. (1940). The New Systematics. Oxford: The Clarendon Press.Google Scholar
Huxley, T. (1862). The anniversary address. Geological Society of London, Quarterly Journal, 18, xl–liv.CrossRefGoogle Scholar
Iaccarino, S. (1985). Mediterranean Miocene and Pliocene planktic foraminifera. In Plankton Stratigraphy, ed. Bolli, H. M., Saunders, J. B. & Perch-Nielsen, K., pp. 283–315. Cambridge: Cambridge University Press, 2 volumes.Google Scholar
Imbrie, J. & Imbrie, K. P. (1979). Ice Ages: Solving the Mystery. London: Macmillan.CrossRefGoogle Scholar
Ingle, J. C. (1973). Summary Comments on Neogene Biostratigraphy, Physical Stratigraphy, and Paleo-Oceanography in the Marginal Northeast Pacific Ocean. Initial Reports of the Deep Sea Drilling Project, 18, 949–960.Google Scholar
Israelsky, M. C. (1949). Oscillation chart. American Association of Petroleum Geologists, Bulletin, 33, 92–98.Google Scholar
Ivany, L. C. (1999). So … now what? Thoughts and ruminations about coordinated stasis. Palaios, 14(4), 4 pp.CrossRefGoogle Scholar
Ivany, L. C. & Schopf, K. M. (1996). New perspectives on faunal stability in the fossil record. Palaeogeography, Palaeoecology, Palaeoclimatology, Special Issue, 127, 359 pp.Google Scholar
Jackson, J. B. C. and Johnson, P. A. (2000). Life in the last few million years. In Deep Time: Paleobiology's Perspective, ed. D. H. Erwin & S. L. Wing, Paleobiology, Supplement to vol. 26(4), 221–235.CrossRef
Jackson, J. B. C., Budd, A. F. & Pandolfi, J. M. (1996). The shifting balance of natural communities? In Evolutionary Paleobiology: Essays in Honor of James W. Valentine, ed. Jablonski, D., Erwin, D. H. & Lipps, J. H., pp. 89–122. Chicago: University of Chicago Press.Google Scholar
Jenkins, D. G. (1960). Planktonic foraminifera from the Lakes Entrance oil shaft, Victoria, Australia. Micropaleontology, 6, 345–371.CrossRefGoogle Scholar
Jenkins, D. G.(1966a). Planktonic foraminiferal zones and new taxa from the Danian to lower Miocene of New Zealand. New Zealand Journal of Geology & Geophysics, 8, 1088–1126.CrossRefGoogle Scholar
Jenkins, D. G.(1966b). Planktonic foraminiferal datum planes in the Pacific and Trinidad Tertiary. New Zealand Journal of Geology & Geophysics, 9, 424–427.CrossRefGoogle Scholar
Jenkins, D. G.(1967). Planktonic foraminiferal zones and new taxa from the Lower Miocene to the Pleistocene of New Zealand. New Zealand Journal of Geology & Geophysics, 10, 1064–1078.CrossRefGoogle Scholar
Jenkins, D. G.(1968). Variations in the numbers of species and subspecies of planktonic Foraminiferida as an indicator of New Zealand Cenozoic paleotemperatures. Paleogeography, Paleoclimatology, Paleoecology, 5, 309–313.CrossRefGoogle Scholar
Jenkins, D. G.(1971). New Zealand Cenozoic planktonic foraminifera. Lower Hutt: New Zealand Geological Survey Paleontological Bulletin, 56, 175 pp.Google Scholar
Jenkins, D. G.(1973). The present status and future progress in the study of Cenozoic planktonic foraminifera. Revista Española de Micropaleontología, 5, 133–146.Google Scholar
Jenkins, D. G.(1985). Southern mid-latitude Paleocene to Holocene planktonic foraminifera. In Plankton Stratigraphy, ed. Bolli, H. M., Saunders, J. B. & Perch-Nielsen, K., pp. 263–282. Cambridge: Cambridge University Press.Google Scholar
Jenkins, D. G.(1993). Cenozoic southern mid- and high-latitude biostratigraphy and chronostratigraphy based on planktonic foraminifera. In The Antarctic Paleoenvironment: a Perspective on Global Change, ed. Kennett, J. P. & Warnke, D. A., pp. 125–144. Washington, D.C.: Antarctic Research Series, Part Two, vol. 60, American Geophysical Union.CrossRefGoogle Scholar
Johnson, D. A. & Nigrini, C. (1985). Time-transgressive late Cenozoic radiolarian events of the equatorial Indo-Pacific. Science, 230, 538–540.CrossRefGoogle ScholarPubMed
JOIDES (Joint Oceanographic Institutes for Deep Earth Sampling) (1981). Report of the Conference on scientific ocean drilling, November 16–18, 1981. Washington: JOI Inc.
Jukes, J. B. (1862). The Student's Manual of Geology. Edinburgh: Adam & Charles Black, 763 pp.Google Scholar
Kauffman, E. G. (1969). Evolutionary rates and biostratigraphy. In North American Paleontological Convention (Chicago, 1969), Proc., Part F (Correlation by Fossils).
Kauffman, E. G.(1977). Evolutionary rates and biostratigraphy. In Concepts and Methods of Biostratigraphy, ed. Kauffman, E. G & Hazel, J. E., pp. 109–142. Stroudsberg, PA: Dowden, Hutchison & Ross.Google Scholar
Kauffman, E. G.(1987). The uniformitarian albatross. Palaios, 2, 531.CrossRefGoogle Scholar
Kauffman, E. G., Elder, W. P. & Sageman, B. B. (1991). High-resolution correlation: a new tool in stratigraphy. In Cycles and Events in Stratigraphy, ed. Einsele, G.et. al., pp. 795–819. Berlin: Springer-Verlag.Google Scholar
Kellogg, T. B (1975). In Investigation of Late Quaternary Paleoceanography and Paleoclimatology, ed. Cline, R. M. & Hays, J. D.. Geological Society of America Memoir 145.Google Scholar
Kelly, D. C., Arnold, A. J. & Parker, W. C. (1996). Paedomorphosis and the origin of the planktonic foraminiferal genus Morozovella. Paleobiology, 22, 266–281.CrossRefGoogle Scholar
Kelly, D. C., Bralower, T. J. & Zachos, J. C. (1998). Evolutionary consequences of the latest Paleocene thermal maximum for tropical planktonic foraminifera. Palaeogeography, Palaeoclimatology, Palaeoecology, 141, 139–161.CrossRefGoogle Scholar
Kemple, W. G., Sadler, P. M. & Strauss, D. J. (1995). Extending graphic correlation to N dimensions. In Graphic Correlation and the Composite Standard Approach, ed. Mann, K. O., Lane, H. R., & Stein, J. R., pp. 65–82. Tulsa: SEPM (Society for Sedimentary Geology), Special Publication No. 53.CrossRefGoogle Scholar
Kennett, J. P. (1968). Latitudinal variation in Globigerina pachyderma (Ehrenberg) in surface sediments of the south-west Pacific Ocean. Micropaleontology, 14, 305–319.CrossRefGoogle Scholar
Kennett, J. P.(1982). Marine Geology. Englewood Cliffs: Prentice-Hall, 813 pp.Google Scholar
Kennett, J. P. & Shackleton, N. J. (1976). Oxygen isotopic evidence for the development of the psychrosphere 38 Myr ago. Nature, 260, 513–515.CrossRefGoogle Scholar
Kennett, J. P. & Stott, L. D. (1991). Abrupt deep-sea warming, paleoceanographic changes and benthic extinctions at the end of the Paleocene. Nature, 353, 225–229.CrossRefGoogle Scholar
Kitcher, P. (1984). Species. Philosophy of Science, 51, 303–333.CrossRefGoogle Scholar
Kitts, D. B. (1966 (1977)). Geologic time. Journal of Geology, 74, 127–146.CrossRefGoogle Scholar
Kitts, D. B.(1977). The Structure of Geology. Dallas: SMU Press.Google Scholar
Kleinpell, R. M. (1979). Criteria in Correlation: Relevant Principles of Science. Bakersfield, California: Pacific Section of the American Association of Petroleum Geologists, 44 pp.Google Scholar
Klitgord, K. D. & Schouten, H. (1986). Plate kinematics of the central Atlantic. In The Geology of North America, Vol. M, The Western North Atlantic Region, ed. Vogt, P. R. & Tucholke, B. E., pp. 351–377. Boulder: Geological Society of America.Google Scholar
Knowlton, N. (1993). Sibling species in the sea. Annual Review of Ecology and Systematics, 24, 189–216.CrossRefGoogle Scholar
Knowlton, N.(2000). Molecular analysis of species boundaries in the sea. Hydrobiologia, 420, 73–90.CrossRefGoogle Scholar
Knox, R. W. O'B. (1996). Correlation of the early Paleogene in northwest Europe: an overview. In Correlation of the Early Paleogene in Northwest Europe, ed. Knox, R. W. O'B., Corfield, R. M. & Dunay, R. E., London: Geological Society, Special Publication 101, 1–11.Google Scholar
Koch, P. L., Zachos, J. C. & Gingerich, P. D. (1992). Correlations between isotope records in marine and continental reservoirs near the Paleocene/Eocene boundary. Nature, 385, 319–322.CrossRefGoogle Scholar
Kominz, M. A., Miller, K. G. & Browning, J. V. (1998). Long-term and short-term global Cenozoic sea-level estimates. Geology, 26, 311–314.2.3.CO;2>CrossRefGoogle Scholar
Kouwenhoven, T. J., Seidenkrantz, M.-S. & Zwaan, G. J. (1999). Deep-water changes: the near-synchronous disappearance of a group of benthic foraminifera from the late Miocene Mediterranean. Palaeogeography, Palaeoclimatology, Palaeoecology, 152, 259–281.CrossRefGoogle Scholar
Krassilov, V. A. (1974). Causal biostratigraphy. Lethaia, 7, 173–179.CrossRefGoogle Scholar
Krassilov, V. A.(1978). Organic evolution and natural stratigraphical classification. Lethaia, 11, 93–104.CrossRefGoogle Scholar
Krijgsman, W., Gaboardi, S., Hilgen, F. J., Iaccarino, S., Kaenel, E. & Laan, E. (2004). Revised astrochronology for the Ain el Beida section (Atlantic Morocco): no glacio-eustatic control for the onset of the Messinian salinity crisis. Stratigraphy, 1, 87–101.Google Scholar
Kucera, M. & Darling, K. F. (2002). Cryptic species of planktonic foraminifera: their effect on palaeoceanographic reconstructions. Philosophical Transactions of the Royal Society. Series A, 360, 695–718.CrossRefGoogle ScholarPubMed
Kucera, M. & Malmgren, B. A. (1998). Differences between evolution of mean form and evolution of new morphotypes: an example from late Cretaceous planktonic foraminifera. Paleobiology, 24, 49–63.CrossRefGoogle Scholar
Laskar, J. (1999). The limits of Earth orbital calculations for geological time-scale use. Philosophical Transactions of the Royal Society of London A, 357, 1735–1760.CrossRefGoogle Scholar
Laskar, J., Joutel, F. & Boudin, F. (1993). Orbital, precessional, and insolation quantities for the Earth from 20 myr to 10 myr. Astronomy & Astrophysics, 270, 522–533.Google Scholar
Laudan, R. (1976). From Mineralogy to Geology: The Foundations of a Science, 1650–1830. Chicago: University of Chicago Press.Google Scholar
Laudan, R.(1982). Tensions in the concept of geology: natural history or natural philosophy? Journal of the History of the Earth Sciences, 1, 7–13.CrossRefGoogle Scholar
Laudan, R.(1987). From Mineralogy to Geology: The Foundations of a Science, 1650–1830. Chicago: University of Chicago Press, 278 pp.CrossRefGoogle Scholar
Laudan, R.(1989). Individuals, species and the development of mineralogy and geology. In What the Philosophy of Biology Is: Essays Dedicated to David Hull, ed. Ruse, M., pp. 221–233. Dordrecht: Kluwer Academic Publishers.CrossRefGoogle Scholar
Lazarus, D. B., (1983). Speciation in pelagic protists and its study in the planktonic microfossil record: a review. Paleobiology, 9, 327–340.CrossRefGoogle Scholar
Lazarus, D. B., Hilbrecht, H., Spencer–Cervato, C. & Thierstein, H. (1995). Sympatric speciation and phylogenetic change in Globorotalia truncatulinoides. Paleobiology, 21, 28–51.CrossRefGoogle Scholar
Lear, C. H., Elderfield, H. & Wilson, P. A. (2000). Cenozoic deep-sea temperatures and global ice volumes from Mg–Ca in benthic foraminiferal calcite. Science, 287, 269–272.CrossRefGoogle ScholarPubMed
Leckie, R. M. (1989). A paleoceanographic model for the early evolutionary history of planktonic foraminifera. Palaeogeography, Palaeoclimatology, Palaeoecology, 73, 107–138.CrossRefGoogle Scholar
Lee, M. S. Y. (2003). The geometric meaning of macroevolution. Trends in Ecology and Evolution, 18, 263–266.CrossRefGoogle Scholar
LeRoy, L. W. (1948). The foraminifer Orbulina universa d'Orbigny, a suggested middle Tertiary time indicator. Journal of Palaeontology, 22, 500–508.Google Scholar
Lethiers, F. (1983). Les extensions stratigraphiques des espèces d'Ostracodes sur les plates-formes dévoniennes: le concept de répartitions sigmoidales. Palaeogeography, Palaeoecology, Palaeoclimatology, 43, 299–312.CrossRefGoogle Scholar
Leupold, W. & van der Vlerk, J. M. (1931). The Tertiary [of the East Indies]. Overdruk uit Leidsche Geol. Mededeelingen, Deel vol. (Festbundel K. Martin), 611–648.
Levinton, J. (2001). Genetics, Paleontology, and Macroevolution, Second Edition. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Li, Q., & McGowran, B. (1994). Miocene upwelling events: foraminiferal evidence from southern Australia. Australian Journal of Earth Sciences, 41, 593–603.CrossRefGoogle Scholar
Li, Q., & McGowran, B.(1997). Miocene climatic oscillations recorded in the Lakes Entrance oil shaft, southeastern Australia: benthic foraminiferal response on a mid-latitude margin. Micropaleontology, 43, 149–164.CrossRefGoogle Scholar
Li, Q., & McGowran, B.(2000). The Miocene foraminifera from Lakes Entrance Oil Shaft, southeastern Australia. Association of Australasian Palaeontologists, Memoir 22, 142 pp., 26 plates.Google Scholar
Li, Q., Davies, P. J., McGowran, B. & Linden, T., (1999). Foraminiferal sequence biostratigraphy of the Oligo-Miocene Janjukian strata from Torquay, southeastern Australia. Australian Journal of Earth Sciences 46, 261–273.CrossRefGoogle Scholar
Li, Q., Davies, P. J., McGowran, B. & Van der Linden, T.(2003). High-resolution foraminiferal ecostratigraphy: an experimental study on upper Oligocene cool-water carbonates from southeastern Australia. In, Olson, H. and Leckie, M. (eds), Paleobiological, Geochemical, and other Proxies of Sea Level Change, pp. 147–171, SEPM Special Volume 75(Society of Sedimentary Geology).Google Scholar
Li, Q., McGowran, B., James, N. P. & Bone, Y., 1996. Foraminiferal biofacies on the mid-latitude Lincoln Shelf, South Australia: oceanographic and sedimentological implications. Marine Geology, 129, 285–312.CrossRefGoogle Scholar
Li, Q., McGowran, B. & White, M. R. (2000). Foraminiferal biostratigraphic and biofacies packages in the mid-Cainozoic Gambier Limestone, South Australia: reappraisal of foraminiferal evidence. Australian Journal of Earth Sciences, 47, 955–970.CrossRefGoogle Scholar
Li, Q., Simo, J. A., McGowran, B. & Holbourn, A., (2004). The eustatic and tectonic origin of Neogene unconformities from the Great Australian Bight. Marine Geology, 203, 57–81.CrossRefGoogle Scholar
Lindsay, E. H. (1989). The setting. In European Neogene Mammal Chronology, ed. Lindsay, E. H., Fahlbusch, V. & Mein, P., pp. 1–15. NATO Advanced Research Workshop on European Neogene Mammal chronology, Schloss Reisensburg. New York & London: Plenum Press.CrossRefGoogle Scholar
Lindsay, E. H. & Tedford, R. H. (1989). Development and application of land mammal ages in North America and Europe, a comparison. In European Neogene Mammal Chronology, ed. Lindsay, E. H., Fahlbusch, V. & Mein, P., pp. 601–624. NATO Advanced Research Workshop on European Neogene Mammal chronology, Schloss Reisensburg. New York & London: Plenum Press.CrossRefGoogle Scholar
Lindsay, E. H., Jacobs, L. L. & Butler, R. F. (1978). Biostratigraphy and magnetostratigraphy of Paleocene terrestrial deposits, San Juan Basin, New Mexico. Geology, 6, 425–429.2.0.CO;2>CrossRefGoogle Scholar
Lindsay, E. H., Opdyke, N. D., Johnson, N. M. & Butler, R. F. (1987). Mammalian chronology and the magnetic polarity time scale. In Cenozoic Mammals of North America: Geochronology and Biostratigraphy, ed. Woodburne, M. O., 269–284. Berkeley: University of California Press.Google Scholar
Lindsay, J. M. (1967). Foraminifera and stratigraphy of the type section of Port Willunga Beds, Aldinga Bay, South Australia. Transactions Royal Society of South Australia, 91, 93–110.Google Scholar
Lindsay, J. M.(1981). Tertiary stratigraphy and foraminifera of the Adelaide City area, St Vincent Basin, South Australia. Unpublished MSc Thesis, The University of Adelaide, Adelaide.
Lindsay, J. M.(1985). Aspects of South Australian foraminiferal biostratigraphy, with emphasis on studies of Massilina and Subbotina. In Stratigraphy, Palaeontology, Malacology, Papers in Honour of Dr Nell Ludbrook, ed. J. M. Lindsay, pp. 187–214. Adelaide Department of Mines and Energy, Special Publication 5, D. J. Woolman, Government Printer.
Lipps, J. H. (1970). Plankton evolution. Evolution, 24, 1–22.CrossRefGoogle ScholarPubMed
Lipps, J. H.(1981). What, if anything, is micropaleontology? Paleobiology. 7, 167–199.CrossRefGoogle Scholar
Lipps, J. H.ed. (1993). Fossil Prokaryotes and Protists. Boston: Blackwell Scientific Publishers.Google Scholar
Loeblich, A. R. & Tappan, H. (1957a). Correlation of the Gulf and Atlantic Coastal Plain Paleocene and Lower Eocene formations by means of planktonic foraminifera. Journal of Paleontology, 31, 1109–1136.Google Scholar
Loeblich, A. R. & Tappan, H. (1957b). Planktonic foraminifera of Paleocene and early Eocene age from the Gulf and Atlantic Coastal Plain. In Loeblich, A. R., & Collaborators (1957), Studies in foraminifera. US National Museum Bulletin 215, 173–198.Google Scholar
Loeblich, A. R. & Tappan, H.(1964a). Sarcodina, chiefly ‘Thecamoebians’ and Foraminiferida, in Treatise on Invertebrate Paleontology, ed. Moore, R. C.. Boulder: Geological Society of America, Part C, vols. 1–2, 900 pp.Google Scholar
Loeblich, A. R. & Tappan, H.(1964b). Foraminiferal classification and evolution. Journal of the Geological Society of India, 5, 5–40.Google Scholar
Loeblich, A. R.et al. (1957). Studies in foraminifera. US National Museum Bulletin 215.Google Scholar
Lohmann, G. P. & Malmgren, B. A. (1983). Equatorward migration of Globorotalia truncatulinoides ecophenotypes through the late Pleistocene:gradual evolution or ocean change? Paleobiology, 9, 414–421.CrossRefGoogle Scholar
Loutit, T. S., ed. (1992). Sea Level Working Group (SLWG), Report. JOIDES Journal, 18, 28–36.
Loutit, T. S. & Kennett, J. P. (1981). New Zealand and Australian Cenozoic cycles and global sea-level changes. The American Association of Petroleum Geologists Bulletin, 65, 1586–1601.Google Scholar
Loutit, T. S., Hardenbol, J. & Vail, P. R. (1988). Condensed sections: the key to age determination and correlation of continental margin sequences. In Sea-Level Changes, an Integrated Approach, ed. Wilgus, C. K., Hastings, B. S., Kendall, C. G. S. C., Posamentier, H. W., Ross, C. A. & Wagoner, J. C.. Society of Economic Paleontologists and Mineralogists, Memoir, 42, 183–213.CrossRefGoogle Scholar
Loutit, T. S., Romine, K. K. & Foster, C. B. (1997). Sequence biostratigraphy, petroleum exploration and A. cinctum. The APPEA Journal 1997, 272–284.Google Scholar
Lovejoy, A. O. (1936). The Great Chain of Being. Cambridge, MA: Harvard University Press.Google Scholar
Lowman, S. W. (1949). Sedimentary facies in the Gulf coast. The American Association of Petroleum Geologists Bulletin, 39, 1939–1947.Google Scholar
Ludbrook, N. H. (1971). Stratigraphy and correlation of marine sediments in the western part of the Gambier Embayment, In The Otway Basin in Southeasten Australia, ed. Wopfner, H., , H. & Douglas, J. G., pp. 4–6. Adelaide and Melbourne, Geological Surveys of South Australia & Victoria, Special Bulletin.Google Scholar
Ludbrook, N. H.(1973). Distribution and Stratigraphic Utility of Cenozoic Molluscan Faunas in Southern Australia. Tohoku University, Science Reports, 2ndseries (Geology), Special Volume No. 6 (Hatai Memorial Volume), pp. 241–261.Google Scholar
Ludbrook, N. H. & Lindsay, J. M. (1969). Tertiary foraminiferal zones in South Australia. In Proceedings of the First International Conference on Planktonic Microfossils, ed. Brönnimann, P. & Renz, H. H., pp. 36–74, vol. 2. Leiden: E. J. Brill.Google Scholar
Luterbacher, H.-P. (1964). Studies in some Globorotalia from the Paleocene and lower Eocene of the central Apennines. Ecologae Geologicae Helvetiae, 57, 631–730.Google Scholar
Luterbacher, H.-P.(1998). Sequence stratigraphy and the limitations of biostratigraphy in the marine Paleogene strata of the Tremp Basin (central part of the southern Pyrenean foreland basins, Spain). In Mesozoic and Cenozoic Sequence Stratigraphy of European Basins, ed. Graciansky, P.-C., Hardenbol, J., Jacquin, T., & Vail, P. R., SEPM (Society of Sedimentary Geology) Special Publication No. 60, 303–310.CrossRefGoogle Scholar
Luterbacher, H.-P. & Premoli Silva, I. (1964). Biostratigrafia del limite Cretaceo–Terziarionell' Appennino centrale. Rivista Italiana Paleontologia Stratigrafia, 70, 67–128.Google Scholar
Luyendyk, B. (1977). Deepsea drilling on the Ninetyeast Ridge: synthesis and a tectonic model. In Indian Ocean Geology and Biostratigraphy, Studies Following Deep Sea Drilling Legs 22–29, ed. Heirtzler, J. R., Bolli, H. M., Davies, T. A., Saunders, J. B., & Sclater, J. G., Washington, D.C.: American Geophysical Union, p. 165–168.Google Scholar
Lyell, C. (1871). Students Elements of Geology. London: John Murray.CrossRefGoogle Scholar
MacFadden, B. J. & Hulburt, R. Jr (1988). Explosive speciation at the base of the adaptive radiation of Miocene grazing horses. Nature, 336, 466–468.CrossRefGoogle Scholar
MacLeod, N. (1991). Punctuated anagenesis and the importance of stratigraphy to paleobiology. Paleobiology, 17, 167–188.CrossRefGoogle Scholar
MacLeod, N.(1993). The Maastrichtian-Danian radiation of triserial and biserial planktic foraminifera: testing phylogenetic and adaptational hypotheses in the (micro) fossil record. Marine Micropaleontology, 21, 547–100.CrossRefGoogle Scholar
MacLeod, N.(2001). The role of phylogeny in quantitative paleobiological data analysis. Paleobiology 27, 226–240.2.0.CO;2>CrossRefGoogle Scholar
MacLeod, N. & Keller, G. (1991a). Hiatus distributions and mass extinctions at the Cretaceous–Tertiary boundary. Geology, 19, 109–147.2.3.CO;2>CrossRefGoogle Scholar
MacLeod, N. & Keller, G.(1991b). How complete are Cretaceous–Tertiary boundary sections? A chronostratigraphic estimate based on graphic correlation. Geological Society of America, Bulletin, 103, 1439–1457.2.3.CO;2>CrossRefGoogle Scholar
MacLeod, N. & Sadler, P. (1995). Estimating the line of correlation. In Graphic Correlation and the Composite Standard Approach, ed. Mann, K. O., Lane, H. R., Stein, J. R., pp. 51–64. Tulsa: SEPM (Society for Sedimentary Geology), Special Publication No. 53.CrossRefGoogle Scholar
Magnus, D. (1996). Heuristics and biases in evolutionary biology. Biology and Philosophy, 12, 1–20.CrossRefGoogle Scholar
Mahner, M. & Bunge, M. (1997). Foundations of Biophilosophy. Berlin & New York: Springer.CrossRefGoogle Scholar
Malmgren, B. A. & Berggren, W. A. (1987). Evolutionary changes in some late Neogene planktonic foraminiferal lineages and their relationship to paleoceanographic changes. Paleoceanography, 2, 445–456.CrossRefGoogle Scholar
Malmgren, B. A. & Kennett, J. P. (1981). Phyletic gradualism in a late Cenozoic planktonic foraminiferal lineage; DSDP Site 284, southwest Pacific. Paleobiology, 7, 230–240.CrossRefGoogle Scholar
Malmgren, B. A., Berggren, W. A. & Lohmann, G. P. (1983). Evidence for punctuated gradualism in the late Neogene Globorotalia tumida lineage of planktonic foraminifera. Paleobiology, 9, 377–388.CrossRefGoogle Scholar
Malmgren, B. A., Berggren, W. A. & Lohmann, G. P.(1984). Species formation through punctuated gradualism in planktonic foraminifera. Science, 225, 317–319.CrossRefGoogle ScholarPubMed
Malmgren, B. A., Kucera, M. & Ekman, G. (1996). Evolutionary changes in supplementary apertural characteristics of the late Neogene Sphaeroidinella dehiscens lineage (planktonic foraminifera). Palaios, 11, 192–206.CrossRefGoogle Scholar
Mancini, E. & Tew, B. H. (1991). Relationships of Paleogene stage and planktonic foraminiferal zone boundaries to lithostratigraphic and allostratigraphic contacts in the eastern Gulf Coastal Plain. Journal of Foraminiferal Research, 21, 48–66.CrossRefGoogle Scholar
Mancini, E. & Tew, B. H.(1995). Geochronology, biostratigraphy and sequence stratigraphy of a marginal marine to marine shelf stratigraphic succession: Upper Paleocene and Lower Eocene, Wilcox Group, eastern Gulf Coastal Plain, U.S.A. In Geochronology Time Scales and Global Stratigraphic Correlation, ed. W. A. Berggren, D. V. Kent, M.-P. Aubry & J. Hardenbol. Tulsa, SEPM Special Publication 54, 28–94, Mann, K. O. & Lane, H. R., eds. (1995). Graphic Correlation. Tulsa: SEPM (Society of Sedimentary Geology) Special Publication, 53.
Margulis, L. & Fester, R. (1991). Symbiosis as a Source of Evolutionary Innovation. Cambridge, MA: MIT Press.Google ScholarPubMed
Marshall, C. R. (1997). Confidence intervals on stratigraphic ranges with nonrandom distributions of fossil horizons. Paleobiology, 23, 165–173.CrossRefGoogle Scholar
Martin, R. E. (1991). Beyond biostratigraphy: micropaleontology in transition? Palaios, 6, 437–438.CrossRefGoogle Scholar
Martin, R. E.(1995). The once and future profession of micropaleontology. Journal of Foraminiferal Research, 25, 372–373.CrossRefGoogle Scholar
Martin, R. E. & Fletcher, R. R. (1995). Graphic correlation of Plio-Pleistocene sequence boundaries, Gulf of Mexico: oxygen isotopes, ice volume, and sea level. In Graphic Correlation and the Composite Standard Approach, ed. Mann, K. O., Lane, H. R. & Stein, J. R., pp. 235–248. Tulsa: SEPM (Society for Sedimentary Geology), Special Publication No. 53.CrossRefGoogle Scholar
Martin, R. E., Neff, E. D., Johnson, G. W. & Krantz, D. E. (1993). Biostratigraphic expression of Pleistocene sequence boundaries, Gulf of Mexico. Palaios, 8, 155–171.CrossRefGoogle Scholar
Marwick, A. (1989). The Nature of History. 3rd edition. Houndmills, Basingstoke: Macmillan, 442 pp.CrossRefGoogle Scholar
Mayr, E. (1942). Systematics and the Origin of Species from the Viewpoint of a Zoologist. New York: Columbia University Press.Google Scholar
Mayr, E.(1961). Cause and effect in biology. Science 134, 1501–1506.CrossRefGoogle ScholarPubMed
Mayr, E.(1963). Animal Species and Evolution. Cambridge: Harvard University Press.CrossRefGoogle Scholar
Mayr, E.(1964). Introduction. In C. Darwin, On the Origin of Species [Facsimile of first edition, 1859, ed. E. Mayr], vii–xxvii. Cambridge, MA: Harvard University Press.Google Scholar
Mayr, E.(1969). Principles of Systematic Zoology. New York: McGraw-Hill, 428 pp.Google Scholar
Mayr, E.(1976). Evolution and the Diversity of Life. Cambridge: Harvard University Press.Google Scholar
Mayr, E.(1982). The Growth of Biological Thought: Diversity, Evolution, and Inheritance. Cambridge, MA: Belknap Press, 974 pp.Google Scholar
Mayr, E.(1988). Toward a New Philosophy of Biology: Observations of an Evolutionist. Cambridge, MA: Belknap Press of Harvard University Press, 564 p.Google Scholar
Mayr, E.(1997). This is Biology: The Science of the Living World. Cambridge, MA: Belknap Press of Harvard University Press, 327 pp.Google Scholar
Mayr, E.(1992). Speciational evolution or punctuated equilibria. In The Dynamics of Evolution, ed. Somit, A. & Peterson, S., pp. 21–48. New York: Cornell University Press.Google Scholar
Mayr, E.(2002). What Evolution Is. London: Phoenix.Google Scholar
Mayr, E. & Ashlock, P. D. (1991). Principles of Systematic Zoology, 2nd Edition. New York: McGraw-Hill, 475 pp.Google Scholar
McDougall, I. & Tarling, D. H. (1963). Dating of the polarity zones in the Hawaiian Islands. Nature, 200, 171–172.CrossRefGoogle Scholar
McGowan, J. A. (1986). The biogeography of pelagic ecosystems. In Pelagic Biogeography, ed. Pierrot–Bults, A. C.et al., 191–200. Paris: UNESCO.Google Scholar
McGowan, J. A. & Walker, P. W. (1993). Pelagic diversity patterns. In Species Diversity in Ecological Communities, ed. Ricklefs, R. E. & Schluter, D., pp. 203–214. Chicago: University of Chicago Press.Google Scholar
McGowran, B. (1968a). Reclassification of Early Tertiary Globorotalia. Micropaleontology, 14, 179–198.CrossRefGoogle Scholar
McGowran, B.(1968b). Late Cretaceous and Early Tertiary correlations in the Indo-Pacific region. In Cretaceous-Tertiary Formations of South India, ed. Rao, L. Rama. Bangalore: Geological Society of India, Memoir 2, 335–360.Google Scholar
McGowran, B.(1971). On foraminiferal taxonomy. In Proceedings of the II Planktonics Conference, Roma 1970, ed. Farinacci, A., 813–820. Roma: Edizioni Technoscienza.Google Scholar
McGowran, B.(1977). Maastrichtian to Eocene foraminiferal assemblages in the northern and eastern Indian Ocean region: correlations and historical patterns. In Indian Ocean Geology and Biostratigraphy, Studies Following Deep Sea Drilling Legs 22–29, ed. Heirtzler, J. R., Bolli, H. M., Davies, T. A., Saunders, J. B. & Sclater, J. G., p. 417–458, Washington, D.C.: American Geophysical Union.Google Scholar
McGowran, B.(1978a). Early Tertiary biostratigraphy in southern Australia: a progress report. In, The Crespin Volume: Essays in Honour of Irene Crespin, ed. Belford, D. J. & Scheibnerova, V.. Canberra: Bureau Mineral Resources Australia, Geology & Geophysics, Bulletin, 192, 83–95.Google Scholar
McGowran, B.(1978b). Stratigraphic record of early Tertiary oceanic and continental events in the Indian Ocean region. Marine Geology, 26, 1–39.CrossRefGoogle Scholar
McGowran, B.(1979). The Australian Tertiary: foraminiferal overview. Marine Micropaleontology, 4, 235–264.CrossRefGoogle Scholar
McGowran, B.(1986a). Beyond classical biostratigraphy. PESA Journal, 9, 28–41.Google Scholar
McGowran, B.(1986b). Cainozoic oceanic events: the Indo-Pacific biostratigraphic record. In Global Stratigraphic Correlation of Mesozoic and Cainozoic Sediments, ed. W. A. Berggren, Palaeogeography, Palaeoclimatology, Palaeoecology, 55, 247–265.
McGowran, B.(1989a). The later Eocene transgressions in southern Australia. Alcheringa, 13, 45–68.CrossRefGoogle Scholar
McGowran, B.(1989b). Silica burp in the Eocene ocean. Geology, 17, 857–860.2.3.CO;2>CrossRefGoogle Scholar
McGowran, B.(1990). Fifty million years ago. American Scientist, 78, 30–39.Google Scholar
McGowran, B.(1991). Evolution and environment in the early Palaeogene. In, The World of Martin F. Glaessner, ed. Radhakrishna, B. P.. Geological Society of India, Memoir 20, 21–53.Google Scholar
McGowran, B., & Beecroft, A. (1985). Guembelitria in the early Tertiary of southern Australia and its palaeoceanographic significance. In Stratigraphy, Palaeontology and Malacology, Papers in Honour of N. H. Ludbrook, ed. Lindsay, J. M.. South Australian Dept Mines & Energy, Special Publication, 5, 247–261.Google Scholar
McGowran, B. & Li, Q. (1994). The Miocene oscillation in southern Australia. Special Volume on Australian Vertebrate Evolution, Palaeontology, & Systematics, Records of the South Australian Museum, 27, 197–212.Google Scholar
McGowran, B. & Li, Q.(1996). Ecostratigraphy and sequence biostratigraphy, with a neritic foraminiferal example from the Miocene in southern Australia. Historical Biology, 11, 137–169.CrossRefGoogle Scholar
McGowran, B. & Li, Q.(2000). Evolutionary palaeoecology of Cainozoic foraminifera: Tethys, IndoPacific, southern Australia. Historical Biology, 15, 3–28.CrossRefGoogle Scholar
McGowran, B. & Li, Q.(2002). Sequence biostratigraphy and evolutionary palaeoecology: Foraminifera in the Cenozoic Era. Memoirs of the Association of Australasian Palaeontologists, 27, 167–188.Google Scholar
McGowran, B., Archer, M., Bock, P. et al. (2000). Australasian palaeobiogeography: the Palaeogene and Neogene record. In Palaeobiogeography of Australasian faunas and floras, ed. Wright, A. J., Young, G. C., Talent, J. A. & Laurie, J. R., pp. 405–470. Australasian Association of Palaeontologists, Memoir, 23.Google Scholar
McGowran, B., Holdgate, G. R., Li, Q. & Gallagher, S. J. (2004). Cenozoic statigraphic succession in southeastern Australia. Australian Journal of Earth Sciences, 51, 459–496.CrossRefGoogle Scholar
McGowran, B., Li, Q., Cann, J.Padley, D., Mckirdy, D. & Shafik, S. (1997b). Biogeographic impact of the Leeuwin Current in southern Australia since the late middle Eocene. Palaeogeography, Palaeoclimatology, Palaeoecology, 136, 19–40.CrossRefGoogle Scholar
McGowran, B. Li, Q. & Moss, G. (1997a). The Cenozoic neritic record in southern Australia: The biogeohistorical framework. In Cool-water Carbonates in Space and Time, ed. James, N. P. & Clarke, J. D. A., Society of Economic Paleontologists & Mineralogists, Special Volume, 56, 185–203.CrossRefGoogle Scholar
McGowran, B., Lindsay, J. M. & Harris, W. K. (1971). Attempted reconciliation of Tertiary biostratigraphic systems, Otway Basin. In The Otway Basin in Southeasten Australia, ed. Wopfner, H. H. & Douglas, J. G., pp. 27–81. Adelaide and Melbourne, Geological Surveys of South Australia & Victoria, Special Bulletin.Google Scholar
McGowran, B., Moss, G. & Beecroft, A. (1992). Late Eocene and early Oligocene in southern Australia: local neritic signals of global oceanic changes. In, Eocene-Oligocene Climatic and Biotic Evolution, ed. Prothero, D. R. & Berggren, W. A., pp. 178–201. Princeton: Princeton University Press.CrossRefGoogle Scholar
McLean, D. M. (1981). Cretaceous–Tertiary extinctions and possible terrestrial and extraterrestrial causes. Syllogeous, 39. Ottawa: National Museums of Canada.Google Scholar
Mein, P. (1975). Résultats du Groupe de Travail des Vertébrés. In Report on Activity of R.C.M.S. Working Group, ed. Senes, J.. Regional Committee of Mediterranean Stratigraphy, 75–81.Google Scholar
Mein, P.(1981). Mammal zonations: introduction. In Annales Géologiques des Pays Helléniques, ed. Marinos, G. & Symeonidis, G., pp. 83–88. Proceeding VIIth International Congress on Mediterranean Neogene, Athens, 1979.Google Scholar
Mein, P.(1989). Die Kleinsäugerfauna des Untermiozäns (Eggenburgien) von Maigen, Niederösterreich. Annales Naturhistorisches Museum Wien, 90A, 49–58.Google Scholar
Miall, A. D. (1997). The Geology of Stratigraphic Sequences: Springer-Verlag, Berlin.CrossRefGoogle Scholar
Miall, A. D.(2004). Empiricism and model building in stratigraphy: the historical roots of present-day practices. Stratigraphy, 1, 3–25.Google Scholar
Miall, A. D. & Miall, C. E. (2001). Sequence stratigraphy as a scientific enterprise: the evolution and persistence of conflicting paradigms. Earth Science Reviews, 54, 321–348.CrossRefGoogle Scholar
Miall, A. D. & Miall, C. E.(2004). Empiricism and model building in stratigraphy: around the hermeneutic circle in the pursuit of stratigraphic correlation. Stratigraphy, 1, 27–46.Google Scholar
Miall, C. E. & Miall, A. D. (2002). The Exxon factor: the roles of corporate and academic science in the emergence and legitimation of a new global model of sequence stratigraphy. The Sociological Quarterly, 43, 307–334.CrossRefGoogle Scholar
Miller, A. I. (1997). Coordinated stasis or coincident relative stability? Paleobiology, 23, 155–164.CrossRefGoogle Scholar
Miller, F. X. (1977). The graphic correlation method in biostratigraphy. In Concepts and Methods of Biostratigraphy, ed. Kauffman, E. G. & Hazel, J. E., pp. 165–186. Stroudsburg: Dowden, Hutchinson & Ross.Google Scholar
Miller, K. G. (1994) The rise and fall of sea level studies: are we at a stillstand? Paleoceanography, 9, 183–184.CrossRefGoogle Scholar
Miller, K. G. & Kent, D. V. (1987). Testing Cenozoic eustatic changes: The critical role of stratigaphic resolution. Cushman Foundation for Foraminiferal Research, Special Publication, 24, 51–55.Google Scholar
Miller, K. G., Fairbanks, R. G. & Mountain, G. S. (1987). Tertiary oxygen isotope synthesis, sea-level history, and continental margin erosion. Paleoceanography, 1, 1–19.CrossRefGoogle Scholar
Miller, K. G., Mountain, G. S., Browning, J. V., Kominz, M., Sugarman, P. J., Christie-Blick, N., Katz, M. E., & Wright, J. D. (1998). Cenozoic global sea level, sequences, and the New Jersey transect: results from coastal plain and continental slope drilling. Reviews of Geophysics, 36, 569–601.CrossRefGoogle Scholar
Miller, K. G., Sugarman, P. J., Browning, J. V., Kominz, M. A., Hernández, J. C., Olsson, R. K., Wright, J. D., Feigenson, M. D. & Sickel, W. V. (2003). Late Cretaceous chronology of large, rapid sea-level changes: glacioeustasy during the greenhouse world. Geology, 31, 585–588.2.0.CO;2>CrossRefGoogle Scholar
Miller, K. G., Thompson, P. R. & Kent, D. V. (1993). Integrated late Eocene–Oligocene stratigraphy of the Alabama coastal plain: correlation of hiatuses and stratal surfaces to glacioeustatic lowerings. Paleoceanography, 8, 313–331.CrossRefGoogle Scholar
Miller, K. G., Wright, J. D. & Fairbanks, R. G. (1991) Unlocking the ice house: Oligocene-Miocene oxygen isotopes, eustasy, and margin erosion. Journal of Geophysical Research, 96(B4), 6829–6848.CrossRefGoogle Scholar
Miller, W. III (1986). Paleoecology of benthic community replacement. Lethaia, 19, 225–231.CrossRefGoogle Scholar
Miller, W. III(1990). Hierarchy, individuality and paleoecosystems. Paleontological Society Special Publication, 5, 31–47.Google Scholar
Miller, W. III(1993). Models of recurrent fossil assemblages. Lethaia, 26, 182–183.CrossRefGoogle Scholar
Mishler, B. & Brandon, R. (1987). Indivuality, pluralism, and the biological species concept. Biology and Philosophy, 2, 397–414.CrossRefGoogle Scholar
Montanari, A., Odin, G. S. & Coccioni, R., Editors (1997). Miocene Stratigraphy: an Integrated Approach. New York: Elsevier.Google Scholar
Monty, C. L. V. (1968). D'Orbigny's concepts of stage and zone. Journal of Paleontology, 42, 689–701.Google Scholar
Moore, R. C. (1941). Stratigraphy. In, Geology, 188–938, 50th Anniversary volume, Geological Society of America, 178–220.Google Scholar
Moore, R. C.(1948). Stratigraphical paleontology. Geological Society of America, Bulletin, 59, 301–326CrossRefGoogle Scholar
Moore, T. C. & Romine, K. (1981). In search of biostratigraphic resolution. In The Deep Sea Drilling Program: a Decade of Progress, ed. Warme, J. E., Douglas, R. G. & Winterer, J. E.. Tulsa: Society of Economic Paleontologists and Mineralogists Special Publication 32, 317–334.CrossRefGoogle Scholar
Morgans, H. E. G., Scott, G. H., Beu, A. G., Graham, I. J., Mumme, T. C., George, W. St. & Strong, C. P. (1996). New Zealand Cenozoic time scale (version 11–96). Institute of Geological & Nuclear Sciences, Science Report 96/38, 12 pp.
Morris, P. J., Ivany, L. C., Schopf, K. M. & Brett, C. E. (1995). The challenge of paleoecological stasis: reassessing sources of evolutionary stability. Proceedings of the National Academy of Sciences USA, 92, 1126–1273.CrossRefGoogle ScholarPubMed
Moss, G. D. & McGowran, B. (2003). Oligocene neritic foraminifera in Southern Australia: spatiotemporal biotic patterns reflect sequence–stratigraphic environmental patterns. In Paleobiological, Geochemical, and Other Proxies of Sea Level Change, ed. Olson, H. C., H. & , R. M.Leckie, M., SEPM (Society of Sedimentary Geology) Special Volume 75, 117–138.Google Scholar
Murphy, M. A. (1977). On time-stratigraphic units. Journal of Paleontology, 51, 213–219.Google Scholar
Murray, G. E. (1961). Geology of the Atlantic and Gulf Coastal Province of North America. New York: Harper & Row.Google Scholar
Murray, J. (1912). The Ocean: A General Account of the Science of the Sea. London: Williams & Norgate.Google Scholar
Nagappa, Y. (1959). Foraminiferal biostratigraphy of the Cretaceous–Eocene succession in the India–Pakistan–Burma region. Micropaleontology, 5, 145–192.CrossRefGoogle Scholar
Nagel, E. (1952). The logic of historical analysis. Science Monthly, 74, 162–169.Google Scholar
Neal, J. E., Stein, J. A. & Gamber, J. A. (1998). Nested stratigraphic cycles and depositional systems of the Paleogene central North Sea. In Mesozoic and Cenozoic Sequence Stratigraphy of European Basins, ed. Graciansky, P.-C., Hardenbol, J., Jacquin, T. & Vail, P. R., SEPM (Society of Sedimentary Geology) Special Publication No. 60, 261–288.CrossRefGoogle Scholar
Norell, M. A. (1992). Taxic origin and temporal diversity: the effect of phylogeny. In Extinction and phylogeny, ed. Novacek, M. J. & Wheeler, Q. D., pp. 89–118. New York: Columbia University Press.Google Scholar
Norell, M. A. & Novacek, M. J. (1992). The fossil record and evolution: comparing cladistic and paleontological evidence for vertebrate history. Science, 255, 1690–1693.CrossRefGoogle Scholar
Norris, R. D.(1991a). Biased extinctions and evolutionary trends. Paleobiology, 17, 388–399.CrossRefGoogle Scholar
Norris, R. D.(1991b). Parallel evolution in the keel structure of planktonic foraminifera. Journal of Foraminiferal Research, 21, 319–331.CrossRefGoogle Scholar
Norris, R. D.(1992). Extinction, selectivity and ecology in planktonic foraminifera. Palaeogeography, Palaeoecology, Palaeoclimatology, 95, 1–17.CrossRefGoogle Scholar
Norris, R. D.(1996). Symbiosis as an evolutionary innovation in the radiation of Paleocene planktonic foraminifera. Paleobiology, 22, 461–480.CrossRefGoogle Scholar
Norris, R. D.(1999). Hydrographic and tectonic control of plankton distribution and evolution. In Reconstructing Ocean History: A Window into the Future, ed. Abrantes, F. & Mix, A., pp. 173–193. London: Plenum.CrossRefGoogle Scholar
Norris, R. D.(2000). Pelagic species diversity, biogeography, and evolution. In Deep time: Paleobiology's perspective, ed. Erwin, D. H & Wing, S. L., Paleobiology, Supplement to vol. 26(4), 236–258.Google Scholar
Norris, R. D. & Vargas, C. (2000). Evolution all at sea. Nature, 405, 23–24.CrossRefGoogle ScholarPubMed
Norris, R. D. & Nishi, H. (2001). Evolutionary trends in coiling of tropical Paleogene planktic foraminifera. Paleobiology, 27, 327–347.2.0.CO;2>CrossRefGoogle Scholar
Norris, R. D. & Röhl, U. (1999). Carbon cycling and chronology of climate warming during the Paleocene/Eocene transition. Nature, 401, 775–778.CrossRefGoogle Scholar
Norris, R. D. & Wilson, P. A. (1998). Low-latitude sea-surface temperatures for the mid-Cretaceous and the evolution of planktic foraminifera. Geology, 26, 823–826.2.3.CO;2>CrossRefGoogle Scholar
Norris, R. D., Corfield, R. M. & Cartlidge, J. E. (1994). Evolutionary ecology of Globorotalia (Globoconella) (planktic foraminifera). Marine Micropaleontology, 23, 121–45.CrossRefGoogle Scholar
Norris, R. D., Corfield, R. M. & Cartlidge, J. E.(1996). What is gradualism? Cryptic species in globorotaliid planktonic foraminifera. Paleobiology, 22, 386–405.CrossRefGoogle Scholar
Odin, G. S. & Curry, D. (1985). Palaeogene time scale: radiometric dating versus magnetostratigraphic approach. Journal Geological Society of London, 142, 1179–1188.CrossRefGoogle Scholar
Odin, G. S., & Curry, D., Gale, N. H. & Kennedy, W. J. (1982). The Phanerozoic time scale in 1981. In Numerical Dating in Stratigraphy, ed. Odin, G. S., pp. 957–960. New York: John Wiley & Sons.Google Scholar
Okada, H. & Bukry, D. (1980). Supplementary modification and introduction of code numbers to the low latitude coccolith biostratigraphic zonation. Marine Micropaleontology, 5, 321–325.CrossRefGoogle Scholar
Oldroyd, D. (1979). Historicism and the rise of historical geology, History of Science, 17, 191–213; 227–257.CrossRefGoogle Scholar
Oldroyd, D.(1996). Thinking about the Earth: A History of Ideas in Geology. London: Athlone, 410 p.Google Scholar
Olson, E. C. (1952). The evolution of a Permian vertebrate chrono-fauna. Evolution, 6, 181–196.CrossRefGoogle Scholar
Olsson, R. K. (1988). Foraminiferal modeling of sea-level change in the late Cretaceous of New Jersey. In Sea-level Changes, An Integrated Approach, ed. Wilgus, C. K., Hastings, B. S., Kendall, C. G. S. C., Posamentier, H. W., Ross, C. A. & Wagoner, J. C., 289–297. Society of Sedimentary Geology, Memoir, 42.CrossRefGoogle Scholar
Olsson, R. K.(1991). Cretaceous to Eocene sea level fluctuations on the New Jersey margin. Sedimentary Geology, 70, 195–208.CrossRefGoogle Scholar
Olsson, R. K. & Wise, W. W. (1987). Upper Paleocene to middle Eocene depositional sequences and hiatuses in the New Jersey Atlantic margin. In Timing and Depositionial History of Eustatic Sequences: Constraits on Seismic Statrigraphy, ed. Ross, C. and Harnan, D.. Cushman Foundation for Foraminiferal Research, Special Publication, 24, 85–97.Google Scholar
Olsson, R. K., Berggren, W. A., Hemleber, C. H. & Huber, B. T. (1999). Atlas of Paleocene Planktonic foraminifera. Washington, DC: Smithsorian Institution Press.Google Scholar
Olsson, R. K., Miller, K. G., Browning, J. V., Wright, J. D. & Cramer, B. S. (2002). Sequence stratigraphy and sea level change across the Cretaceous-Tertiary boundary on the New Jersey passive margin. In Catastropic events and mass extinctions: impacts and beyond. Geological Society of America Special Paper, 56, 97–108.Google Scholar
Orue-Etxebarria, X., Pujalte, V., Bernaola, G., Apellaniz, E., Baceta, J. I., Payros, A., Nuñez-Betelu, , Serra-Kiel, J. & Tosquella, J. (2001). Did the Late Paleocene thermal maximum affect the evolution of large foraminifers? Evidence from the Campo section (Pyrenees, Spain). Marine Micropaleontology, 41, 45–71.CrossRefGoogle Scholar
Osborn, H. F. (1910). The Age of Mammals in Europe, Asia and North America. New York: Macmillan, 635 pp.CrossRefGoogle Scholar
Osborn, H. F.(1934). Aristogenesis, the creative principle in the origin of species. American Naturalist, 68, 193–235.CrossRefGoogle Scholar
Oslick, J. S., Miller, K. G., Feigenson, M. D. & Wright, J. D. (1994). Oligocene–Miocene strontium isotopes: stratigraphic revisions and correlations to an inferred glacioeustatic record. Paleoceanography, 9, 427–444.CrossRefGoogle Scholar
Padian, K., Lindberg, D. R. & Polly, P. D. (1994). Cladistics and the fossil record: the uses of history. Annual Review of Earth and Planetary Sciences, 22, 63–69.CrossRefGoogle Scholar
Paley, W. (1827). The Works of William Paley, D. D., Archdeacon of Carlisle. Edinburgh: Peter Brown and Thomas Nelson.Google Scholar
Pälike, H. & Shackleton, N. J. (2000). Constraints on astronomical parameters from the geological record for the last 25 Myr. Earth and Planetary Science Letters, 182, 1–14.CrossRefGoogle Scholar
Pandolfi, J. M. (1996). Limited membership in Pleistocene reef coral assemblages from the Huon Peninsula, Papua New Guinea: constancy during global change. Paleobiology, 22, 152–176.CrossRefGoogle Scholar
Papp, A. & Schmid, M. E. (1985). Die Fossilen Foraminiferen des Tertiären Beckens von Wien: Revision der Monographie von Alcide d'Orbigny (1846). Abhandlungen der Geologischen Bundesanstalt, Band 37.Google Scholar
Papp, A., Grill, R., Janoschek, R., Kapounek, J, Kollmann, K.Turnovsky, K. (1968). Zur Nomenklatur des Neogens in Österreich. Sonderabdruck aus den Verhandlungen der Geologischen Bundesanstalt 1968, Heft 1/2, 9–18 [English text by E. J. Tynan, p. 19–27].Google Scholar
Park, J., D'Hondt, S. L., King, J. W. & Gibson, C. (1993). Late Cretaceous precessional cycles in double time: a warm-earth Milankovitch response. Science, 261, 1431–1434.CrossRefGoogle ScholarPubMed
Pawlowski, J. (2000). Introduction to the molecular systematics of foraminifera. In Advances in the Biology of Foraminifera, ed. Lee, J. J. & Hallock, P., p. 1–12. Micropaleontology, 46, Supplement 1.Google Scholar
Pearson, P. N. (1998a). Stable isotopes and the study of evolution in planktonic foraminifera. Paleontological Society Papers, 4, 138–178.Google Scholar
Pearson, P. N.(1998b). Evolutionary concepts in biostratigraphy. In Unlocking the Stratigraphical Record, ed. Doyle, P. & Bennett, M. R., pp. 123–144. John Wiley & Sons, Chichester.Google Scholar
Pearson, P. N.(1998c). Speciation and extinction asymmetries in paleontological phylogenies: evidence for evolutionary progress? Paleobiology, 24: 305–335.Google Scholar
Pearson, P. N., Olsson, R. K., Huber, B. T., Hemleben, C. & Berggren, W. A. (eds.) (in press) Atlas of Eocene Planktonic Foraminifera, Cushman Foundation for Foraminiferal Research, Special Publication.
Pearson, P. N., Shackleton, N. J. & Hall, M. A. (1997). Stable isotope evidence for the sympatric divergence of Globigerinoides trilobus and Orbulina universa (planktonic foraminifera). Journal of the Geological Society, 154, 295–302.CrossRefGoogle Scholar
Pessagno, E. A. (1967). Upper Cretaceous planktonic foraminifera from the western Gulf Coastal Plain. Paleontographica Americana, 5, 259–444.Google Scholar
Phillips, J. (1840). Palaeozoic Series. In The Penny Cyclopedia (cited in Berry, 1968).
Phillips, J.(1861). Life on the Earth: its Origin and Succession. Cambridge & London.Google Scholar
Phillips, J.(1829). Organic remains of the eastern part of Yorkshire. In Conkin & Conkin (1984), pp. 86–93.
Pimm, A. C., McGowran, B. & Gartner, S. (1974). Early sinking history of Ninetyeast Ridge, north-eastern Indian Ocean. Bulletin Geological Society America, 85, 1219–1224.2.0.CO;2>CrossRefGoogle Scholar
Poag, C. W. (1977). Biostratigraphy in Gulf Coast petroleum exploration. In Concepts and Methods of Biostratigraphy, ed. Kauffman, E. G. & Hazel, J. E., pp. 213–234. Stroudsberg, PA: Dowden, Hutchison & Ross.Google Scholar
Pokorny, V. (1963). Principles of Zoological Micropalaeontology. Translated by K. A. Allen. Oxford: Pergamon.Google Scholar
Pomerol, Ch. & Premoli Silva, I., Editors (1986). Terminal Eocene Events. New York: Elsevier Science.Google Scholar
Popper, K. R. (1957). The Poverty of Historicism. London: Routledge & Kegan Paul.Google Scholar
Prell, W. L., Imbrie, J., Morley, J. J., Pisias, N. G., Shackleton, N. J. & Streeter, H. F. (1986). Graphic correlation of oxygen isotope stratigraphy: applications to the late Quaternary. Paleoceanography, 1, 137–162.CrossRefGoogle Scholar
Premoli Silva, I., Coccioni, R. & Montanari, A., Editors (1988). The Eocene–Oligocene boundary in the Marche–Umbria basin (Italy). International Subcommission on Paleogene Stratigraphy, International Union of Geological Sciences, Eocene–Oligocene meeting, Ancona (Italy), Special Publication.Google Scholar
Press, F. & Siever, R. (1978). Earth, 2nd Edition. San Francisco: W. H. Freeman.Google Scholar
Prothero, D. R. (1992). Punctuated equilibrium at twenty: a paleontological perspective. Skeptic, 1, 38–47.Google Scholar
Prothero, D. R.(1994a). The late Eocene–Oligocene extinctions. Annual Reviews of Earth & Planetary Science, 22, 145–165.CrossRefGoogle Scholar
Prothero, D. R.(1994b). The Eocene–Oligocene Transition: Paradise Lost. New York: Columbia University Press, 291 pp.Google Scholar
Prothero, D. R.(1995). Geochronology and magnetostratigraphy of Paleogene North American land mammal ‘ages’: an update. In Geochronology Time Scales and Global Stratigraphic Correlation, ed. Berggren, W. A., Kent, D. V., Aubry, M.-P. & Hardenbol, J.. Tulsa: SEPM (Society of Sedimentary Geology) Special Publication, 54, 305–316.CrossRefGoogle Scholar
Prothero, D. R.(1998). The chronological, climatic, and paleogeographic background to North American mammalian evolution. In Evolution of the Tertiary Mammals of North America, Volume 1, ed. Janis, C. M, Scott, K. M. & Jacobs, L. L., pp. 9–36. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Prothero, D. R.(1999). Does climatic change drive mammalian evolution? GSA Today, 9, 1–7.Google Scholar
Prothero, D. R. & Berggren, W. A., Editors (1992). Eocene–Oligocene Climatic and Biotic Evolution. Princeton: Princeton University Press, 568 pp.CrossRefGoogle Scholar
Prothero, D. R. & Heaton, T. H. (1996). Faunal stability during the early Oligocene climatic crash. Palaeogeography, Palaeoecology, Palaeoclimatology, 127, 257–283.CrossRefGoogle Scholar
Quillévéré, F., Norris, R. D., Moussa, I. & Berggren, W. A. (2001). Role of photosymbiosis and biogeography in the diversification of early Paleogene acarininids (planktonic foraminifera). Paleobiology 27, 311–326.2.0.CO;2>CrossRefGoogle Scholar
Rabeder, G. (1986). Herkunft und frühe Evolution der Gattung Microtus (Arvicolidae, Rodentia). Sonderdruck aus Z. f. Säugetierkunde, Bd.51, H.6, 350–367.
Raffi, I. (1999). Precision and accuracy of nannofossil biostratigraphic correlation. Philosophical Transactions of the Royal Society of London A, 357: 1975–1994.CrossRefGoogle Scholar
Rama Rao, L. (1968). The problem of the Cretacous–Tertiary boundary. In Cretaceous–Tertiary Formations of South India, ed. Rao, L. Rama. Bangalore: Geological Society of India, Memoir 2, 1–9.Google Scholar
Raup, D. M. (1991). Extinction: Bad Genes or Bad Luck? New York: W. W. Norton.Google ScholarPubMed
Raup, D. M. & Sepkoski, J. J. Jr (1984). Periodicity of extinctions in the geologic past. Proceeding of the National Academy of Sciences USA, 81, 801–805.CrossRefGoogle ScholarPubMed
Raup, D. M. & Stanley, S. M. (1978). Principles of Paleontology. New York, San Francisco: W. H. Freeman, 481 pp.Google Scholar
Ravelo, A. C. & Fairbanks, R. G. (1992). Oxygen isotopic composition of multiple species of planktonic foraminifera: recorders of the modern photic zone temperature gradient. Paleoceanography, 7, 815–831.CrossRefGoogle Scholar
Raymo, M. (1994) The initiation of northern hemisphere glaciation. Annual Reviews of Earth and Planetary Science, 22, 353–384.CrossRefGoogle Scholar
Rea, D. K., Zachos, J. C., Owen, R. M. & Gingerich, P. D. (1990). Global change at the Paleocene/Eocene boundary: climatic and evolutionary consequences of tectonic events. Palaeogeography, Palaeoclimatology, Palaeoecology, 79, 117–128.CrossRefGoogle Scholar
Reeckmann, A. (1994). Geology of the onshore Torquay Sub-basin: a sequence-stratigraphic approach. In Otway Basin Symposium, Extended Abstracts, compiled D. M. Finlayson. Melbourne: Australian Geological Survey Organisation, Record 1994/13, pp. 3–6.Google Scholar
Reif, W.-E. (1993). Afterword. In Schindewolf (1993), pp. 435–453.
Reiss, Z. (1968). Planktonic foraminiferids, stratotypes, and a reappraisal of Neogene chronostratigraphy in Israel. Israel Journal of Earth Sciences, 17, 153–169.Google Scholar
Remane, J. (1997). Chronostratigraphic standards: how are they defined and when should they be changed? Quaternary International, 40, 3–4.CrossRefGoogle Scholar
Remane, J., Bassett, M. G., Cowie, J. W., Gohrbandt, K. H., Lane, H. R., Michelsen, O. & Naiwen, W. (1996). Revised guidelines for the establishment of global stratigraphic standards by the International Commission of Stratigraphy (ICS). Episodes, 18, 77–81.Google Scholar
Rensch, B. (1983). The abandonment of Lamarckian explanations: the case of climatic parallelism of animal characteristics. In Dimensions of Darwinism, ed. Grene, M., 31–42. Cambridge: Cambridge University Press.Google Scholar
Ridley, M. (1986). Evolution and Classification: The Reformation of Cladism. London, New York: Longman, 201 pp.Google Scholar
Riedel, W. R. (1973). Cenozoic planktonic micropaleontology and biostratigraphy. Annual Review of Earth and Planetary Sciences, 1, 241–268.CrossRefGoogle Scholar
Rieppel, O. (1988). Fundamentals of Comparative Biology. Basel, Boston: Birkhauser Verlag, 202 pp.Google Scholar
Rio, D., Sprovieri, R. & Thunell, R. (1991). Pliocene–Pleistocene chronostratigraphy: a reevaluation of Mediterranean type sections. Geological Society of America Bulletin, 103, 1049–1058.2.3.CO;2>CrossRefGoogle Scholar
ROCC Research on Cretaceous Cycles Group (1986). Rhythmic bedding in Upper Cretaceous pelagic carbonate sequences: varying sedimentary response to climatic forcing. Geology, 14, 153–156.2.0.CO;2>CrossRef
Rodgers, J. (1959). The meaning of correlation. American Journal of Science, 257, 684–691.CrossRefGoogle Scholar
Rögl, F. (1985). Late Oligocene and Miocene planktic foraminifera of the central Paratethys. In Plankton Stratigraphy, ed. Bolli, H. M., Saunders, J. B. & Perch-Nielsen, K., pp. 315–328. Cambridge: Cambridge University Press.Google Scholar
Rögl, F.(1998). Palaeogeographic considerations for Mediterranean and Paratethys seaways (Oligocene to Miocene). Annalen des Naturhistorischen Museums in Wien, 99A, 279–310.Google Scholar
Rögl, F.(1999). Mediterranean and Paratethys. Facts and hypotheses of an Oligocene to Miocene paleogeography (short overview). Geologica Carpathica, 50, 339–349.Google Scholar
Rögl, F. & Steininger, F. F. (1983). Vom Zerfall der Tethys zu Mediterran und Paratethys. Die neogene Paläogeographie und Palinspastik des zirkummediterranen Raumes. Ann. Nat. Hist. Mus. Wien 85A, 135–163.Google Scholar
Rögl, F. & Steininger, F. F.(1984). Neogene Paratethys, Mediterranean and Indo–Pacific seaways. Implications for the palaeobiogeography of marine and terrestrial biotas. In Fossils and Climate, ed. Brenchley, P., pp. 171–200. New York: Wiley.Google Scholar
Rohling, E. J. & Cooke, S. (1999). Stable oxygen and carbon isotopes in foraminiferal carbonate shells. In Modern Foraminifera, ed. Gupta, B. K. Sen, pp. 239–259. Dordrecht, Kluwer Academic Publishers.Google Scholar
Rosenberg, A. (1985). The Structure of Biological Science. New York: Cambridge University Press.CrossRefGoogle Scholar
Ross, C. A. & Ross, J. P. (1985). Late Paleozoic depositional sequences are synchronous and worldwide. Geology, 13, 194–197.2.0.CO;2>CrossRefGoogle Scholar
Ruddiman, W. F., Raymo, M. & McIntyre, A. (1986). Matuyana 41,000 year cycles: North Atlantic Ocean and northern hemisphere ice sheets. Earth and Planetary Science Letters, 80, 117–29.CrossRefGoogle Scholar
Rudwick, M. J. S. (1972). The Meaning of Fossils: Episodes in the History of Palaeontology. London, New York: Macdonald and Co., American Elsevier (2nd Edition, 1985).Google Scholar
Rudwick, M. J. S.(1982a). Cognitive styles in geology. In Essays in the Sociology of Perception, ed. Douglas, M., pp. 210–242. London: Routledge, Kegan, Paul.Google Scholar
Rudwick, M. J. S.(1982b). Charles Lyell's dream of a statistical palaeontology. Palaeontology, 21, 225–244.Google Scholar
Rudwick, M. J. S.(1990). Introduction. Introduction to Facsimile of C. Lyell, Principles of Geology, First Edition, London: J. Murray, 1830–1833. pp. vii–viii Chicago: University of Chicago Press.Google Scholar
Rudwick, M. J. S.(1998). Lyell and the Principles of Geology. In Lyell: The Past is the Key to the Present, ed. Blundell, D. J. & Scott, A. C.. Geological Society Special Publication No. 143, 3–15.Google Scholar
Ruse, M. (1986). Taking Darwin Seriously: A Naturalistic Approach to Philosophy. New York: Blackwell.Google Scholar
Ruse, M.(1998). All my love is towards individuals. Evolution, 52, 283–288.CrossRefGoogle Scholar
Sadler, P. M. (2004). Quantitative biostratigraphy – achieving finer resolution in global correlation. Annual Reviews of Earth and Planetary Sciences, 32, 187–213.CrossRefGoogle Scholar
Sadler, P. M. & Cooper, R. A. (2003). Best-fit intervals and consensus sequences: comparison of the resolving power of traditional biostratigraphy and computer– assisted correlation. In High Resolution Approaches to Stratigraphic Paleontology, ed. Harries, P., pp. 49–94. Dordrecht: Kluwer Academic Publishers.Google Scholar
Saito, T. (1977). Late Cenozoic planktonic foraminiferal datum levels: the present state of knowledge toward accomplishing Pan-Pacific correlations. In First International Conference on Pacific Neogene Stratigraphy, Tokyo, 1976, Proceedings, pp. 61–80.Google Scholar
Saito, T.(1984). Planktonic foraminiferal datum planes for biostratigraphic correlation of Pacific Neogene sequences – 1982 status report. In Pacific Neogene Datum Planes: Contributions to Biostratigraphy and Chronology, ed. Ikebe, N. & Tsuchi, R., pp. 47–67. Tokyo: University of Tokyo Press.Google Scholar
Salthe, S. N. (1985). Evolving Hierarchical Systems: Their Structure and Representation. New York: Columbia University Press, 343 pp.Google Scholar
Salvador, A., Editor (1994). International Stratigraphic Guide: A Guide to Stratigraphic Classification, Terminology, and Procedure, Second Edition. International Union of Geological Sciences & Geological Society of America, 214 pp.Google Scholar
Savage, D. E. (1977). Aspects of vertebrate paleontological stratigraphy and geochronology. In Concepts and Methods of Biostratigraphy, ed. Kauffman, E. G. & Hazel, J. E., pp. 427–442. Stroudsberg, PA: Dowden, Hutchison & Ross.Google Scholar
Savage, D. E. & Russell, D. E. (1983). Mammalian Paleofaunas of the World. Reading, MA: Addison-Wesley.Google Scholar
Savin, S. M., Douglas, R. G. & Stehli, F. G. (1975). Tertiary marine paleotemperatures. Geological Society of America, Bulletin, 86, 1499–1510.Google Scholar
Schaeffer, B., Hecht, M. K. & Eldredge, N. (1972). Phylogeny and paleontology. In Evolutionary Biology, ed. Dobzhansky, Th.et al., pp. 31–46. New York: Appleton-Century-Crofts.CrossRefGoogle Scholar
Schenck, H. G. & Muller, S. W. (1941). Stratigraphic terminology. Geological Society of America, Bulletin, 52, 1419–1426.CrossRefGoogle Scholar
Schindewolf, O. H. (1993). Basic Questions in Paleontology: Geologic Time, Organic Evolution, and Biological Systematics. Translated by Judith Schaefer; edited by Reif, Wolf-Ernst. Chicago: University of Chicago Press. Originally published (1950) as Grundfragen der Paläontologie. Stuttgart: E. Schweizerbart'sche Verlagsbuchhandlung.Google Scholar
Schneider, C. & Kennett, J. P. (1996). Isotopic evidence for interspecies habitat differences during evolution of the Neogene planktonic foraminiferal clade Globoconella. Paleobiology, 22, 282–303.CrossRefGoogle Scholar
Schoch, R. M. (1989). Stratigraphy: Principles and Methods. New York: Van Nostrand Reinhold, 375 pp.Google Scholar
Schwan, W. (1980). Geodynamic peaks in alpinotype orogenies and changes in ocean–floor spreading during Late Jurassic–Late Tertiary time. American Association of Petroleum Geologists Bulletin, 64, 359–373.Google Scholar
Schwarzacher, W. & Fischer, A. G. (1982). Limestone–shale bedding and perturbations of the Earth's orbit. In Cyclic and Event Stratification, ed. Einsele, G. & Seilacher, A., pp. 72–95. Berlin/Heidelberg/New York: Springer-Verlag.CrossRefGoogle Scholar
Scott, G. H. (1960). The type locality concept in time-stratigraphy. New Zealand Journal of Geology & Geophysics, 3, 580–584.CrossRefGoogle Scholar
Scott, G. H.(1985). Homotaxy and biostratigraphical theory. Palaeontology, 28, 777–782.Google Scholar
Seibold, E. & Berger, W. H. (1993). The Sea Floor: an Introduction to Marine Geology (2nd ed.). Berlin: Springer Verlag.CrossRefGoogle Scholar
Sepkoski, J. J., (1978). A kinetic model of Phanerozoic taxonomic diversity. I. Analysis of marine orders. Paleobiology, 5, 223–251.CrossRefGoogle Scholar
Sepkoski, J. J., Jr. (1986). Global bioevents and the question of periodicity. In Global Bio-events,ed. Walliser, O. H., pp. 47–61. Berlin, Springer-Verlag.CrossRefGoogle Scholar
Sereno, P. C. (1997). The origin and evolution of dinosaurs. Annual Reviews of Earth and Planetary Sciences, 25, 435–490.CrossRefGoogle Scholar
Sereno, P. C.(1999). The evolution of dinosaurs. Science, 284, 2137–2147.CrossRefGoogle ScholarPubMed
Shackleton, N. J. (1985). Oceanic carbon isotope constraints on oxygen and carbon dioxide in the atmosphere. In The Carbon Cycle and Atmospheric CO2: Natural Variations Archaean to Present, ed. Sundquist, E. T.Broecker, W. S., pp. 412–417. Washington, D.C., American Geophysical Union, Geophysical Monograph 32.CrossRefGoogle Scholar
Shackleton, N. J.(1986). Paleogene stable isotope events. Palaeogeography, Palaeoclimatology, Palaeoecology, 57, 91–102.CrossRefGoogle Scholar
Shackleton, N. J., Crowhurst, S. J., Weedon, G. P. & Laskar, J. (1999). Astronomical calibration of Oligocene–Miocene time, Philosophical Transactions of the Royal Society of London A, 357: 1907–1929.CrossRefGoogle Scholar
Shackleton, N. J. & Kennett, J. P. (1975). Paleotemperature history of the Cenozoic and the initiation of Antarctic glaciation: oxygen and carbon isotope analyses in DSDP sites 277, 279, and 281. In Initial Reports of the Deep Sea Drilling Project, Kennett, J. P, Houtz, R. E, et al. Washingto, D.C.: US Government Printing Office.CrossRefGoogle Scholar
Shackleton, N. J. & Opdyke, N. D. (1973). Oxygen isotope and paleomagnetic stratigraphy of equatorial Pacific core V28–38: oxygene isotope temperatures and ice volumes on a 105 and 106 year scale. Quaternary Research, 3, 39–55.CrossRefGoogle Scholar
Shackleton, N. J., Hall, M. A., Raffi, I., Tauxe, L. & Zachos, J. (2000). Astronomical calibration age for the Oligocene–Miocene boundary. Geology, 28: 447–450.2.0.CO;2>CrossRefGoogle Scholar
Shaw, A. B. (1964). Time in Stratigraphy. New York: McGraw-Hill, 365 pp.Google Scholar
Shaw, A. B.(1969). Adam and Eve, paleontology, and the non-objective arts. Journal of Paleontology, 43, 1085–1093.Google Scholar
Sheehan, P. M., (1996). A new look at ecological evolutionary units (EEUs). Palaeogeography, Palaeoecology, Palaeoclimatology, 127, 21–32.CrossRefGoogle Scholar
Silver, L. T. & Schultz, P. H. (1982). Geological implications of impacts of large asteroids and comets on the earth. Geological Society of America, Special Paper 190.Google Scholar
Simmons, M. D. (1998). Biostratigraphy – surviving extinction. Palaios, 13(3) Online, 2 pp.CrossRefGoogle Scholar
Simmons, M. D., Berggren, W. A., O'Neill, B. J., Scott, R. W., Steininger, F. F. & Ziegler, W. (2000). Biostratigraphy and geochronology. In Fossils and the Future: Paleontology in the 21st Century, ed. R. H. Lane, F. F Steininger, R. L. Kaesler, W. Ziegler & J. H. Lipps. Senckenberg–Buch Nr. 74, 119–132.
Simmons, M. D. & Williams, C. L. (1992). Sequence stratigraphy and eustatic sea-level change: the role of micropalaeontology. Journal of Micropalaeontology, 11, 112.CrossRefGoogle Scholar
Simons, A. M., (2002). The continuity of macroevolution and microevolution. Journal of Evolutionary Biology, 15, 688–701.CrossRefGoogle Scholar
Simpson, G. G. (1944). Tempo and Mode in Evolution. New York: Columbia University Press.Google Scholar
Simpson, G. G.(1949; revised edition, 1967). The Meaning of Evolution. New Haven: Yale University Press.Google Scholar
Simpson, G. G.(1951). The species concept. Evolution, 5, 285–298.CrossRefGoogle Scholar
Simpson, G. G.(1952). Periodicity in vertebrate evolution. Journal of Paleontology, 26, 359–370.Google Scholar
Simpson, G. G.(1953). The Major Features of Evolution. New York: Columbia University Press, 434 pp.Google Scholar
Simpson, G. G.(1961). Principles of Animal Taxonomy. New York: Columbia University Press.Google Scholar
Simpson, G. G.(1964). This View of Life. New York: Harcourt, Brace & World.Google Scholar
Simpson, G. G.(1965). The Geography of Evolution. New York: Capricorn.Google Scholar
Simpson, G. G. (1970). Uniformitarianism: an inquiry into principle, theory, and method in geohistory and biohistory. In Essays in Evolution and Genetics, ed. Hecht, M. K. & Steere, W. C., pp. 43–96. New York: Appleton-Century-Crofts.Google Scholar
Singleton, O. P. (1968). Otway region. In A Regional Guide to Victorian Geology. Ed. McAndrew, J. & Marsden, M. A. H., pp. 117–131. Melbourne: Geology Department, University of Melbourne.Google Scholar
Sloss, L. L. (1963). Sequences in the cratonic interior of North America. Geological Society of America Bulletin, 100, 1661–1665.2.3.CO;2>CrossRefGoogle Scholar
Sloss, L. L., Krumbein, W. C. & Dapples, E. C. (1949). Integrated facies analysis. Geological Society of America Memoir, 39, 91–124.CrossRefGoogle Scholar
Smith, A. B. (1994). Systemmatics and the Fossil Record: Documenting Evolutionary Patterns. Oxford: Blackwell Sctientific Publication.CrossRefGoogle Scholar
Spencer-Cervato, C., Lazarus, D. B., Beckmann, J.-P., Salis Perch-Nielsen, K. & Biolzi, M. (1993). New calibration of Neogene radiolarian events in the North Pacific. Marine Micropaleontology, 21, 261–293.CrossRefGoogle Scholar
Spencer-Cervato, C., Thierstein, H. R., Lazarus, D. B. & Beckmann, J.-P. (1994). How synchronous are Neogene marine plankton events? Paleoceanography, 9, 739–763.CrossRefGoogle Scholar
Spero, H. J. (1992). Do planktic foraminifera accurately recordshifts in the carbon isotopic composition of sea water ΣCO2? Marine Micropaleontology, 19, 275–285.CrossRefGoogle Scholar
Spero, H. J. & DeNiro, M. J. (1987). The influence of symbiont photosynthesis on the δ18O and δ13C values of planktonic foraminiferal shell calcite. Symbiosis, 4, 213–228.Google Scholar
Srinivasan, M. S. & Kennett, J. P. (1976). Evolution and phenotypic variation in the Late Cenozoic Neogloboquadrina dutertrei plexus. In Progress in Micropaleontology, Selected Papers in Honour of Prof. Kiyoshi Asano, ed. Takayanagi, Y. and Saito, T., 329–354. New York: Micropaleontology Press.Google Scholar
Srinivasan, M. S. & Kennett, J. P.(1981a). A review of Neogene planktonic foraminiferal biostratigraphy: applications in the equatorial and south Pacific. In The Deep Sea Drilling Program: a Decade of Progress, ed. Warme, J. E., Douglas, R. G. & Winterer, J. E.. Tulsa: Society of Economic Paleontologists and Mineralogists Special Publication 32, 395–432.CrossRefGoogle Scholar
Srinivasan, M. S. & Kennett, J. P.(1981b). Neogene planktonic foraminiferal biostratigraphy and evolution: equatorial to Subantarctic South Pacific. Marine Micropaleontology, 6, 499–533.CrossRefGoogle Scholar
Srinivasan, M. S. & Sinha, D. K. (1991). Improved correlation of the Late Neogene planktonic foraminiferal datums in the equatorial to cool subtropical DSDP sites, southwest Pacific: application of the graphic correlation method. In The World of Martin F. Glaessner, ed. Radhakrishna, B. P.. Geological Society of India, Memoir 20, 55–94.Google Scholar
Stainforth, R. M., Lamb, J. M., Luterbacher, H., Beard, J. H. & Jeffords, R. M. (1975). Cenozoic Planktonic Foraminifera Zonation and Characteristics of Index Forms. Lawrence, University of Kansas Paleontological Contributions62, 1–245 (in two parts).Google Scholar
Stanley, S. M. (1979). Macroevolution, Pattern and Process. Baltimore: Johns Hopkins University Press, 332 pp.Google Scholar
Stanley, S. M., Wetmore, K. L. & Kennett, J. P. (1988). Macroevolutionary differences between the two major clades of Neogene planktonic foraminifera. Paleobiology 14, 235–249.CrossRefGoogle Scholar
Steininger, F. F. (1977). Integrated assemblage-zone biostratigraphy at marine– non-marine boundaries: examples from the Neogene of central Europe. In Concepts and Methods of Biostratigraphy, ed. Kauffman, E. G. & Hazel, J. E., pp. 235–256. Stroudsburg, PA: Dowden, Hutchinson & Ross, Inc.Google Scholar
Steininger, F. F. & Papp, A. (1979). Current biostratigraphic and radiometric correlations of Late Miocene Central Paratethys stages (Sarmatian s.str., Pannonian s.str., and Pontian) and Mediterranean stages (Tortonian and Messinian) and the Messinian event in the Paratethys. Newsletters in Stratigraphy, 8, 1000–110.CrossRefGoogle Scholar
Steininger, F. F. & Proponents of Working Group (1994). Proposal for the Global Stratotype Section and Point (GSSP) for the base of the Neogene (the Paleogene–Neogene boundary). IUGS/ICS/SNS/Working Group for the Paleogene–Neogene Boundary. Vienna: Institute of Paleontology, University of Vienna, pp. 1–41.Google Scholar
Steininger, F. F., Aubry, M.-P., Berggren, W. A., et al. (1997). The global stratotype and point (GSSP) for the base of the Neogene. Episodes, 2, 23–28.Google Scholar
Steininger, F. F., Berggren, W. A., Kent, D. V., Bernor, R. C., Sen, S., and Agusti, J. (1995). Circum-Mediterranean (Miocene and Pliocene) marine–continental chronologic correlations of European mammal units and zones. In The Evolution of Western Eurasian Neogene Mammal Faunas, ed. Bernor, R. L., Fahlbusch, V., Mittmann, V. & Rietschel, S., pp. 23–46, New York: Columbia University Press.Google Scholar
Steininger, F. F., Rögl, F., Hochuli, P. & Müller, C. (1989). Lignite deposition and marine cycles: the Austrian Tertiary lignite deposits – a case history. Öst. Akad. Wiss., Math.-nat. Kl., Abt.I, Bd 5:309–332.Google Scholar
Steininger, F. F., Rögl, F. & Martini, E. (1976). Current Oligocene/Miocene biostratigraphic concept of the Central Paratethys (middle Europe). Newsletters in Stratigraphy, 4, 174–202.CrossRefGoogle Scholar
Steininger, F. F., Senes, I., Kleemann, K.Rögl, F. (1985). Neogene of the Mediterranean, Tethys and Paratethys, 2 Vols, Wien: Universität Wien, Paläontologisches Institut.Google Scholar
Stott L. D. & Kennett J. P. (1990). Antarctic Paleogene planktonic foraminifer biostratigraphy: ODP Leg 113, Sites 689 and 690. In Proceedings of the Ocean Drilling Program, Scientific Results, 113, Barker, P. F., Kennett, J. P.et al., pp. 549–569, College Station, Texas.Google Scholar
Strasser, A., Hillgärtner, H., Hug, W. & Pittet, B. (2000). Third-order depositional sequences reflecting Milankovitch cyclicity. Terra Nova, 12, 303–311.CrossRefGoogle Scholar
Strasser, A., Pittet, B., Hillgärtner, H., Hug, W. & Pasquier, J.-B. (1999). Depositional sequences in shallow carbonate-dominated sedimentary systems: concepts for a high-resolution analysis. Sedimentary Geology, 128, 201–221.CrossRefGoogle Scholar
Subbotina, N. N. (1953). [Globigerinidae, Hantkeninidae, and Globorotaliidae. Fossil foraminifera of the U.S.S.R.] Vses. Neft. Nauchno–Issled. Geol.-Razved. Inst. (VNIGRI), Trudy, n.s. 6, 9, 1–296, 41 pls.
Sylvester–Bradley, P. C., (1951). The subspecies in palaeontology. Geological Magazine, 88, 88–102.CrossRefGoogle Scholar
Tan Sin Hok (1932). On the genus Cycloclypeus Carpenter. I. Wetensch. Meded. 19. Dienst van der Mijnbouw.
Hok, Tan Sin(1936). Zur Kenntnis der Miogypsiniden. De Ingenieur in Nederlandsch–Indië, No. 3(IV), 45–61, 84–98, 109–123.Google Scholar
Hok, Tan Sin(1937). Weitere Untersuchungen über die Miogypsiniden. De Ingenieur in Nederlandsch-Indië, No. 4(IV), 35–45, 87–111.Google Scholar
Hok, Tan Sin(1939a). On Polylepidina, Orbitocyclina and Lepidorbitoides. De Ingenieur in Nederlandsch–Indië, No. 6(5), 53–83.Google Scholar
Hok, Tan Sin(1939b). The results of phylomorphogenetic studies of some larger Foraminifera (a review). Mijnbouw en Geologie, 6, 93–97.Google Scholar
Hok, Tan Sin(1939c). Remarks on the ‘letter classification’ of the East Indian Tertiary. De Ingenieur in Nederlandsch–Indië, 6 (7), 98–101.Google Scholar
Tang, C. M. & Bottjer, D. J. (1996). Long-term faunal stasis without evolutionary coordination: Jurassic marine benthic paleocommunities, Western Interior, United states. Geology, 24, 815–818.2.3.CO;2>CrossRefGoogle Scholar
Tappan, H. & Lipps, J. H. (1966). Wall structures, classification, and evolution in planktonic foraminifera. American Association of Petroleum Geologists, Bulletin, 50, 637.Google Scholar
Tedford, R. H. (1970). Principles and practices of mammalian geochronology in North America. In, North American Paleontological Convention (Chicago, 1969), Proc., Part F (Correlation by Fossils), 666–703.
Teggart, F. J. (1925). Theory of History. New Haven: Yale University Press.Google Scholar
Teichert, C., (1958). Some biostratigraphical concepts. Geological Society of America Bulletin, 69, 99–120.CrossRefGoogle Scholar
Templeton, A. (1989). The meaning of species and speciation: a genetic perspective. In Otte & Endler.
Ten Kate, W. G. H. & Sprenger, A. (1993). Orbital cyclicities above and below the cretaceous–Paleogene boundary at Zumaya (N. Spain), Agost and Relleu (SE Spain). Sedimentary Geology, 87, 69–101.CrossRefGoogle Scholar
Thaler, L. (1965). Une échelle de zones biochronologiques pour les mammifères du Tertiaire d'Europe. C. R. Somm. S.G.P.F, 1965, 118.
Thaler, L.(1966). Les rongeurs fossiles du Bas Languedoc dans leur rapport avec l'histoire des faunes et la stratigraphie du Tertiaire d'Europe. Mém. Mus. Nat. Hist. Natur., Nouvelle Séries., Séries C, 17, 1–295.Google Scholar
Thalmann, H. E. (1934). Die regional–stratigraphische Verbreitung der oberkretazischen Foraminiferen–Gattung Globotruncana Cushman 1927. Ecologae Geologicae Helvetiae, 27, 413–428.Google Scholar
Thierstein, H. R., Geitzenauer, K., Molfino, B. & Shackleton, N. J. (1977). Global synchronicity of Late Quaternary coccolith datums: validation by oxygen isotopes. Geology, 5, 400–404.2.0.CO;2>CrossRefGoogle Scholar
Thomas, E. (1992). Middle Eocene-late Oligocene bathyal benthic foraminifera (Weddell Sea): faunal changes and implications for ocean circulation. In Eocene–Oligocene Climatic and Biotic Evolution, ed. Prothero, D. R. & Berggren, W. A., pp. 245–271. Princeton: Princeton University Press.CrossRefGoogle Scholar
Thomas, E.(1999). Introduction to ‘Biotic responses to major paleoceanographic changes’. In Reconstructing Ocean History: A Window into the Future, ed. Abrantes, F. & Mix, A., pp. 163–171. London: Plenum.CrossRefGoogle Scholar
Thomas, E., Zachos, J. C. & Bralower, T. J. 2000. Deep-sea environments on a warm earth: latest Paleocene-early Eocene. In Warm Climates in Earth History, ed. Huber, B. T., MacLeod, K. G. & Wing, S. L., 132–160. Cambridge University Press, Cambridge.Google Scholar
Tian, J., Wang, P., Cheng, X. & Li, Q. (2002). Astronomically tuned Plio-Pleistocene benthic δ18O record from South China Sea and Atlantic–Pacific comparison. Earth and Planetary Science Letters, 203, 1015–1029.CrossRefGoogle Scholar
Tjalsma, R. C. & Lohmann, G. P. (1983). Paleocene–Eocene Bathyal and Abyssal Benthic Foraminifera from the Atlantic Ocean. New York: Micropaleontology, Special Paper No. 4.Google Scholar
Toulmin, L. D. (1977). Stratigraphic distribution of Paleocene and Eocene fossils in the eastern Gulf Coast regions. Alabama Geological Survey Monograph 13, vol. 1, 602 pp.Google Scholar
Toumarkine, M. & Luterbacher H.-P. (1985). Paleocene and Eocene planktic foraminifera. In Plankton Stratigraphy, ed. Bolli, H. M., Saunders, J. B. & Perch-Nielsen, K., pp. 87–154. Cambridge, Cambridge University Press, 2 volumes.Google Scholar
Van Couvering, J. A. & Berggren, W. A. (1977). Biostratigraphical basis of the Neogene time scale. In Concepts and Methods of Biostratigraphy, ed. Kauffman, E. G. & Hazel, J. E., 283–306. Stroudsburg: Dowden, Hutchinson & Ross, Inc.Google Scholar
Trümpy, R. (1973). The timing of orogenic events in the central Alps. In Gravity and tectonics, ed. Jong, K. A. & Scholten, R., pp. 229–251. New York: John Wiley.Google Scholar
Vail, P. R., Audemard, F., Bowman, S. A., Eisner, P. N. & Perez–Cruz, C. (1991). The stratigraphic signatures of tectonics, eustasy and sedimentology – an overview. In Cycles and Events in Stratigraphy, ed. Einsele, G., Ricken, W. & Seilacher, A., pp. 617–628. Berlin: Springer-Verlag.Google Scholar
Vail, P. R., Mitchum Jr, R. M., Todd, R. G., Widmier, J. M., Thompson III, S., Sangree, J. B., Bubb, J. N. & Hatlelid, W. G. (1977). Seismic stratigraphy and global changes of sea-level. In Seismic Stratigraphy: Applications to Hydrocarbon Exploration, ed. Payton, C. E. American Association of Petroleum Geologists Memoir, 26, 49–212.Google Scholar
Vakarcs, G., Hardenbol, J., Abreu, V. A., Vail, P. R., Várnai, P. & Tari, G. (1998). Oligocene–Middle Miocene depositional sequences of the Central Paratethys and their correlation with regiona stages. In Mesozoic and Cenozoic Sequence Stratigraphy of European Basins, ed. Graciansky, P.-C., Hardenbol, J., Jacquin, T. & Vail, P. R., SEPM (Society of Sedimentary Geology) Special Publication No. 60, 209–232.CrossRefGoogle Scholar
Valentine, J. W. & May, C. L. (1996). Hierarchies in biology and paleontology. Paleobiology, 22, 23–33.CrossRefGoogle Scholar
Bemmelen, R. W. (1949). The Geology of Indonesia. The Hague: Government Printing Office, 3 Volumes.Google Scholar
Couvering, J. A., Editor (1997). The Pleistocene Boundary and the Beginning of the Quaternary. Cambridge: Cambridge University Press.Google Scholar
Couvering, J. A., Castradori, D., Cita, M. B., Hilgen, F. J. & Rio, D. (2000). The base of the Zanclean Stage and of the Pliocene Series. Episodes, 23, 179–187.Google Scholar
Van Couvering, J. A.; Hedberg, H. D. (1977). Review and response to review of: Hedberg, H. D., Editor, 1976, International Stratigraphic Guide: A Guide to Stratigraphic Classification, Terminology and Procedure, New York: John Wiley & Sons. Micropaleontology, 23, 227–232.Google Scholar
Vlerk, I. M. (1955). Correlation of the Tertiary of the Far East and Europe. Micropaleontology, 1, 72–75.CrossRefGoogle Scholar
Vlerk, I. M. (1959). Problems and principles of Tertiary and Quaternary stratigraphy. Quarterly Journal Geological Society of London, 115, 49–63.CrossRefGoogle Scholar
Vlerk, I. M. & Umbgrove, J. H. F. (1927). Tertiaire gidsforaminiferenvan Nederlandsch Oost–Indië. Dutch East Indies, Dienst. Mijnb., Wetensch. Meded., No. 9.Google Scholar
Van Harten, D. (1988). Chronoecology, a non-taxonomic application of ostracods. In Ostracoda in the Earth Sciences, ed. DeDeckker, P., Colin, J.-P. & Peypouquet, J.-P., p. 47–54. Amsterdam: Elsevier.Google Scholar
Harten, D. & Hinte, J. E, 1984. Ostracod range charts as a chronoecologic tool. Marine Micropaleontology, 8, 425–433.CrossRefGoogle Scholar
van Hinte., J. E. (1969). The nature of biostratigraphic zones. In Proceedings of the First International Conference on Planktonic Microfossils, ed. Brönnimann, P. & Renz, H. H.. Leiden: E. J. Brill, vol. 2, pp. 267–272.Google Scholar
Vaughan, T. W. (1924). American and European Tertiary larger foraminifera. Bulletin Geological Society of America, 35, 785–822.CrossRefGoogle Scholar
Vella, P. (1965). Sedimentary cycles, correlation, and stratigraphic classification. Royal Society of New Zealand, Geology, Transactions, 3, 1–9.Google Scholar
Vénec-Peyré, M.-T. (2004). Beyond frontiers and time: the scientific and cultural heritage of Alcide d'Orbigny (1802–1857). Marine Micropaleontology, 50, 149–159.CrossRefGoogle Scholar
Vincent, E. & Berger, W. H. (1985). Carbon dioxide and polar cooling in the Miocene: the Monterey hypothesis. In The Carbon Cycle and Atmospheric CO2: Natural Variations Archaean to Present, ed. Sundquist, E. T. & Broecker, W. S. W. S., pp. 455–468. Washington, D.C., American Geophysical Union, Geophysical Monograph 32.CrossRefGoogle Scholar
Visser, W. A. & Hermes, J. J., Compilers (1962). Geological Results of the Exploration for Oil in Netherlands New Guinea. The Hague: Geologische serie, Nederlands Geologisch Mijnbouwkundig Genootschap, Special nummer, 265 pp.Google Scholar
Engelhardt, W. (1982). Neptunismus und Plutonismus. Fortschritte der Mineralogie, 60, 21–43.Google Scholar
Vrba, E. S. (1980). Evolution, species and fossils: how does life evolve? South African Journal of Science, 76, 61–84.Google Scholar
Vrba, E. S.(1985). Environment and evolution: alternative causes of the temporal distribution of evolutionary events. South African Journal of Science, 81, 229–236.Google Scholar
Vrba, E. S.(1995). The fossil record of African antelopes (Mammalia, Bovidae) in relation to human evolution and paleoclimate. In Paleoclimate and Evolution, With Emphasis on Human Origins, ed. Vrba, E. S., Denton, G. H., Partridge, T. C. & Burckle, L. H., pp. 385–424. New Haven: Yale University Press.Google Scholar
Vrba, E., Denton, G. H., Partridge, T. H. & Burckle, L. H., Editors (1995). Environmental Change and Evolution. Yale University Press, New HavenGoogle Scholar
Wade, M. (1964). Application of lineage concept to biostratigraphic zoning based on planktonic foraminifera. Micropaleontology, 10, 273–290.CrossRefGoogle Scholar
Wade, M.(1966). Lineages of planktonic foraminifera in Australia. In Committee on Mediterranean Stratigraphy, ed. Drooger, C. W., Reiss, Z., Rutsch, R. F. & Marks, P., pp. 30–39. Proceedings of the third session in Berne, 8–13 June 1964. International Union of Geological Sciences Commission on Stratigraphy, Leiden: E. J. Brill.Google Scholar
Wagner, P. J. (1995). Stratigraphic tests of cladistic hypotheses. Paleobiology, 21, 153–178.CrossRefGoogle Scholar
Wagner, P. J.(2000). Phylogenetic analyses and the fossil record: tests and inferences, hypotheses and models. In Deep Time: Paleobiologys Perspective, ed. D. H. Erwin & S. L. Wing. Paleobiology, Supplement to vol. 26(4), 341–371.CrossRef
Wagner, P. J. & Erwin, D. H. (1995). Phylogenetic patterns as tests of speciation models. In New Approaches to Speciation in the Fossil Record, ed. Erwin, D. H. & Anstey, R. L.. pp. 87–122, Columbia University Press: New York.Google Scholar
Wakefield, M. I. & Monteil, E. (2002). Biosequence stratigraphical and palaeoenvironmental findings from the Cretaceous through tertiary succession, Central Indus Basin, Pakistan. Journal of Micropalaeontology, 21, 115–130.CrossRefGoogle Scholar
Walsh, S. L. (1998). Fossil datum and paleobiological event terms, paleontostratigraphy, chronostratigraphy, and the definition of land mammal ‘age’ boundaries. Journal of Vertebrate Paleontology, 18, 150–179.CrossRefGoogle Scholar
Webb, S. D. (1984). On two kinds of rapid faunal turnover. In Catastrophes and Earth History, ed. Berggren, W. A. & Couvering, J. A., pp. 417–436. Princeton: Princeton University Press.CrossRefGoogle Scholar
Webb, S. D. & Opdyke, N. D. (1995). Global climatic influence on Cenozoic land mammal faunas. In Effects of Past Global Change on Life, ed. Kennett, J. P. & Stanley, S. M., pp. 184–208. Washington, D.C.: National Academy of Sciences, Studies in Geophysics.Google Scholar
Weedon, G. P. (1993). The recognition and implications of orbital forcing of climatic and sedimentary cycles. In Sedimentology Review, ed. Wright, V. P., pp. 31–50. oxford: Blackwell.CrossRefGoogle Scholar
Weedon, G. P.Shackleton, N. J.Pearson, P. N. (1997). The Oligocene time scale and cyclostratigraphy on the Ceara Rise, western equatorial Atlantic. Proceedings of the Ocean Drilling Program, Scientific Results, 154, 101–116. College Station, Texas.Google Scholar
Wei, K.-Y. (1994). Allometric heterochrony in the Pliocene–Pleistocene planktonic foraminiferal clade Globoconella. Paleobiology, 20, 66–84.CrossRefGoogle Scholar
Wei, K.-Y. & Kennett, J. P. (1983). Nonconstant extinction rates of Neogene planktonic foraminifera. Nature, 305, 218–220.CrossRefGoogle Scholar
Wei, K.-Y. & Kennett, J. P.(1988). Phyletic gradualism and punctuated equilibrium in the late Neogene planktonic foraminiferal clade Globoconella. Paleobiology, 14, 345–363.CrossRefGoogle Scholar
Wheeler, H. E. (1958). Time-stratigraphy. American Association of Petroleum Geologists Bulletin, 42, 1047–1063.Google Scholar
Wiley, E. O. (1981). Phylogenetics: The Theory and Practice of Phylogenetic Systematics. New York: Wiley, 439 pp.Google Scholar
Wilson, D. (1993). Confirmation of the astronomical calibration of the magnetic polarity timescale from seafloor spreading rates. Nature, 364, 788–790.CrossRefGoogle Scholar
Wilson, E. O. (1998). Consilience: The Unity of Knowledge. New York: Knopf.Google Scholar
Wilson, R. C. L. (1998). Sequence stratigraphy: a revolution without a cause. In Lyell: The Past is the Key to the Present, ed. Blundell, D. J. & Scott, A. C.. Geological Society Special Publication No. 143, 303–314.Google Scholar
Winchester, S. (2001). The Map that Changed the World: The Tale of William Smith and the Birth of a Science. London: Viking.Google Scholar
Wing, S. L. & Tiffney, B. H. (1987). The reciprocal interaction of angiosperm evolution and tetrapod herbivory. Review of Palaeobotany and Palynology, 50, 179–210.CrossRefGoogle Scholar
Wing, S. L., Sues, H.-D.et al. (1992). Mesozoic and early Cenozoic terrestrial ecosystems. In Terrestrial Ecosystems Through Time: Evolutionary Paleoecology of Terrestrial Plants and Animals, ed. Behrensmeyer, A. K., Damuth, J. D., DiMichele, W. A., Potts, R., Sues, H.-D. & Wing, S. L., The Evolution of Terrestrial Ecosystems Consortium, pp. 327–416. Chicago & London: The University of Chicago Press.Google Scholar
Wolfe, J. A. (1978). A paleobotanical interpretation of Tertiary climates in the Northern Hemisphere. American Scientist, 66, 694–703.Google Scholar
Wood, H. E. II, Chaney, R. W. Jr, Clark, J., Colbert, E. H., Jepsen, G. W., Reeside, J. B., Stock, C. (1941). Nomenclature and correlation of the North American continental Tertiary. Geological Society of America Bulletin, 52, 1–48.CrossRefGoogle Scholar
Woodburne, M. O. (1977). Definition and characterization of mammalian chronostratigraphy. Journal of Paleontology, 51, 220–234.Google Scholar
Woodburne, M. O.ed. (1987). Cenozoic Mammals of North America: Geochronology and Biostratigraphy. Berkeley: University of California Press.Google Scholar
Woodburne, M. & Swisher, C. (1995). Mammal high–resolution geochronology, intercontinental overland dispersals, sea level, climate and vicariance. In Geochronology Time Scales and Global Stratigraphic Correlation, ed. Berggren, W. A., Kent, D. V., Aubry, M.-P.Hardenbol, J.. Tulsa, SEPM (Society of Sedimentary Geology) Special Publication 54, 335–364.CrossRefGoogle Scholar
Woodburne, M. O., Tedford, R. H., Archer, M., Turnbull, W., Plane, M. & Lundelius, E. L. (1985). Biochronology of the continental mammal record of Australia and New Guinea. In Stratigraphy, Palaeontology, Malacology, Papers in Honour of Dr Nell Ludbrook, ed. Lindsay, J. M., pp. 347–365. Adelaide Department of Mines and Energy, Special Publication 5, D. J. Woolman, Government Printer.Google Scholar
Woodruff, F. & Savin, S. M. (1991). Mid-Miocene isotope stratigraphy in the deep sea: high-resolution correlations, paleoclimatic cycles, and sedimentary preservation. Paleoceanography, 6, 755–801.CrossRefGoogle Scholar
Worsley, T. R. & Jorgens, M. L. (1977). Automated biostratigraphy. In Oceanic Micropalaeontology, ed. Ramsay, A. T. S., p. 1201–1229. London: Academic Press.Google Scholar
Wright, J. D & Miller, K. G. (1993). Southern Ocean influences on late Eocene to Miocene deepwater circulation. In The Antarctic Paleonviroment: a Perspective on Global Change, ed. Kennett, J. P. & Warnke, D. A., pp. 125–144. Washington, D.C.: Antarctic Research Series, Part Two, v. 60, American Geophysical Union.CrossRefGoogle Scholar
Wright, J. D., Miller, K. G. & Fairbanks, R. G. (1992). Early and Middle Miocene stable isotopes: implications for deepwater circulation and climate. Paleoceanography, 7, 357–389.CrossRefGoogle Scholar
Young, K. (1960). Biostratigraphy and the new paleontology. Journal of Paleontology, 34, 347–358.Google Scholar
Zachos, J., Arthur, M. A. & Dean, W. E. (1989). Geochemical evidence for the suppression of pelagic marine productivity at the Cretaceous–Tertiary boundary. Nature, 337, 61–64.CrossRefGoogle Scholar
Zachos, J. C., Pagani, M., Sloan, L., Thomas, E. & Billups, K. (2001b). Trends, rythyms, and aberrations in global climate 65 Ma to Present. Science, 292, 686.CrossRefGoogle Scholar
Zachos, J. C., Quinn, T. C. & Salamy, K. C. (1996). High-resolution (104 years) deep-sea stable isotope records of the Eocene–Oligocene transition. Paleoceanography, 11, 251–266.CrossRefGoogle Scholar
Zachos, J. C., Shackleton, N. J., Revenaugh, J. S., Pälike, H. & Flower, B. P. (2001a). Periodic and non–periodic climate response to orbital forcing across the Oligocene–Miocene boundary. Science, 292, 274–277.CrossRefGoogle Scholar
Zalasiewiecz, J., Smith, A., Brenchley, P., Evans, J., Knox, R., Riley, N., Gale, A., Gregory, F. J., Rushton, A., Gibbard, P., Hesselbo, S., Marshall, J., Oates, M., Rawson, P. & Trewin, N. (2004). Simplifying the stratigraphy of time. Geology, 32, 1–4.CrossRefGoogle Scholar
Zinsmeister, W. J. & Feldmann, R. M. (1984). Cenozoic high latitude heterochroneity of southern hemisphere marine faunas. Science, 224, 281–283.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • References
  • Brian McGowran, University of Adelaide
  • Book: Biostratigraphy
  • Online publication: 02 December 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511610653.010
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • References
  • Brian McGowran, University of Adelaide
  • Book: Biostratigraphy
  • Online publication: 02 December 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511610653.010
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • References
  • Brian McGowran, University of Adelaide
  • Book: Biostratigraphy
  • Online publication: 02 December 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511610653.010
Available formats
×