Hostname: page-component-848d4c4894-75dct Total loading time: 0 Render date: 2024-05-02T03:08:33.210Z Has data issue: false hasContentIssue false

The Climate and Landscape of the Middle Part of the Weichselian Glaciation in Europe: The Stage 3 Project

Published online by Cambridge University Press:  20 January 2017

Tjeerd H. van Andel*
Affiliation:
Godwin Institute of Quaternary Research and Department of Earth Sciences, University of Cambridge, Downing Street, Cambridge, CB2 3EQ, United Kingdom, E-mail: vanandel@esc.cam.ac.uk

Abstract

Oxygen isotope stage 3 (OIS 3) was a mild interval between the two cold maxima of the last (Weichselian) glaciation marked by climate changes oscillating on a 100–1000 yr time scale between near-interglacial and peak-glacial conditions. During OIS 3, modern humans entered Europe, and somewhat later their Neanderthal predecessors became extinct. Our understanding of this momentous event depends on an answer to the question, Did the unstable environmental conditions of the time play a significant role in early human history? The Stage 3 Project is an interdisciplinary study with two main goals: (1) to describe with existing data and to simulate the climates and landscapes of typical warm and cold phases between 45,000 and 30,000 yr ago and (2) to compare the results with the spatial and temporal distribution of human beings in this context. This paper introduces the Stage 3 Project and provides background to a set of papers on the climate and landscape aspects of the Project that will appear in Quaternary Research and to studies of their relevance to the Early Upper Paleolithic of Europe to appear in journals yet to be determined.

Type
Research Article
Copyright
University of Washington

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

Ammerman, A.J., and Cavalli-Sforza, L.L. Measuring the rate of spread of early farming in Europe. Man 6, (1971). 674688.CrossRefGoogle Scholar
Ammerman, A.L., and Cavalli-Sforza, L.L. The Neolithic Transition and the Genetics of Population in Europe. (1984). Princeton Univ. Press, Princeton.Google Scholar
Andrén, T., Björck, J., and Johnsen, S. Correlation of Swedish glacial varves with the Greenland (GRIP) oxygen isotope record. Journal of Quaternary Science 14, (1999). 361372.3.0.CO;2-R>CrossRefGoogle Scholar
Bond, G., Showers, W., Cheseby, M., Lotti, R., Almasi, P., Demenocal, P., Priore, P., Cullen, H., Hajdas, I., and Bonani, G. A pervasive millennial-scale cycle in North Atlantic Holocene and glacial climates. Science 278, (1997). 12571266.Google Scholar
Coope, G.R. Fossil coleopteran assemblages as sensitive indicators of climatic change during the Devensian (Last) cold stage. Philosophical Transactions, The Royal Society B 280, (1977). 313340.Google Scholar
Coope, G.R., Gibbard, P.L., Hall, A.R., Preece, R.C., Robinson, J.E., and Sutcliffe, A.J. Climatic and environmental reconstructions based on fossil assemblages from Middle Devensian (Weichselian) deposits of the River Thames at South Kensington, central London. Quaternary Science Reviews 16, (1997). 11631196.CrossRefGoogle Scholar
Davies, W., Stewart, J., and van Andel, T.H. Neanderthal landscapes—A preview. Stringer, C., Barton, R.N.E., and Finlayson, J.C. Neanderthals on the Edge. (2000). Oxbow Books, Oxford. 18.Google Scholar
Gamble, C.S. The Palaeolithic Societies of Europe. (1999). Cambridge Univ. Press, Cambridge.Google Scholar
Grootes, P.M., Stuiver, M., White, J.W., Johnsen, S., and Jouzel, J. Comparison of oxygen isotope records from the GISP2 and GRIP Greenland ice cores. Nature 366, (1993). 552554.Google Scholar
Guiot, J., de Beaulieu, J.L., Cheddadi, R., David, F., Ponel, P., and Reille, M. The climate in western Europe during the last glacial/interglacial cycle derived from pollen and insect remains. Palaeogeography, Palaeoclimatology, Palaeoecology 103, (1993). 7394.Google Scholar
Guthrie, R.D. Frozen Fauna of the Mammoth Steppe: The Story of Blue Babe. (1990). Univ. of Chicago Press, Chicago.Google Scholar
Haxeltine, A., and Prentice, I.C. BIOME3: An equilibrium terrestrial biosphere model based on ecophysiological constraints, resource availability and competition among plant functional types. Global Biogeochemical Cycles 10, (1996). 603628.CrossRefGoogle Scholar
Holdaway, S., and Porch, N. Cyclical patterns in the Pleistocene human occupation of Southwest Tasmania. Archaeology in Oceania 30, (1995). 7482.CrossRefGoogle Scholar
Huntley, B., Watts, W., Allen, J.R.M., and Zolitschka, B. Palaeoclimate, chronology and vegetation history of the Weichselian late glacial: Comparative analysis of data from three cores at Lago Grande di Monticchio (Italy): Initial results. Quaternary Science Reviews 18, (1999). 945960.Google Scholar
Jöris, O., and Weninger, B. Extension of the 14-C calibration curve to ca. 40,000 cal BC by synchronizing Greenland 18O/16O ice core records and North Atlantic Foraminifera profiles: A comparison with U/Th coral data. Radiocarbon 40, (1998). 495504.CrossRefGoogle Scholar
Jöris, O., and Weninger, B. Radiocarbon calibration and the absolute chronology of the Late Glacial. L'Europe Centrale et Septentrionale au Tardiglaciaire, Table Ronde de Nemours 13–16 mai 1997. 7, (2000). 1954.Google Scholar
Kozlowski, J.K. Cultural context of the last Neandertals and early modern humans in central-eastern Europe. XIII International Congress of Prehistoric and Protohistoric Sciences, Forli, Italy, Colloquium 5, (1996). 205218.Google Scholar
Kurtén, B. Pleistocene Mammals of Europe. (1968). Weidenfeld and Nicholson, Google Scholar
Laj, C., Mazaud, A., and Duplessy, J.C. Geomagnetic intensity and 14C abundance in the atmosphere and ocean during the past 50 kyr. Geophysical Research Letters 23, (1996). 20452048.CrossRefGoogle Scholar
Lister, A.M. Giant deer and giant red deer from Kent's Cavern and the status of Strongyloceros spelaeus Owen. Transactions and Proceedings, Torquay Natural History Society 91, (1987). 189198.Google Scholar
Lister, A.M., and Sher, A.V. Ice cores and Mammoth extinction. Nature 378, (1995). 2324.Google Scholar
Martinson, D., Pisias, N.G., Hays, J.D., Imbrie, J., Moore, T.C. Jr., and Shackleton, N.J. Age dating and the orbital theory of the Ice Ages: Development of a high-resolution 0–300,000 year chronostratigraphy. Quaternary Research 27, (1987). 129.Google Scholar
Meese, D.A., Gow, A.J., Alley, R.B., Zielinski, G.A., Grootes, P.M., Ram, M., Taylor, K.C., Mayewski, P.A., and Bolzan, J.F. The Greenland ice-sheet Project 2 depth-age scale: Methods and results. Journal of Geophysical Research 102, (1997). 2641126423.Google Scholar
Mellars, P.A. The Neanderthal Legacy: An Archaeological Perspective from Western Europe. (1996). Princeton Univ. Press, Princeton.Google Scholar
Miskovsky, J.-C. Les applications de la géologie à la reconnaissance de l'environnement de l'homme fossile. Mémoires de la Société Géologique de France 160, (1992). Google Scholar
Porter, S.C. Some geological implications of average Quaternary glacial conditions. Quaternary Research 32, (1989). 245261.CrossRefGoogle Scholar
Roebroeks, W., Conard, N.J., and van Kolfschoten, T. Dense forests, cold steppes and the Palaeolithic settlement of northern Europe. Current Anthropology 33, (1992). 551586.Google Scholar
Semino, O., Passarino, G., Oefner, P.J., Lin, A.A., Arbuzova, S., Beckman, L.E., de Benedictis, G., Francalacci, P., Kouvatsi, A., Limborska, S., Marcikiae, M., Mika, A., Mika, B., Primorac, D., Santachiara-Benerecetti, A.S., Cavalli-Sforza, L.L., and Underhill, P.A. The genetic legacy of Paleolithic Homo sapiens sapiens in extant Europeans: A Y-chromosome perspective. Science 290, (2000). 11551159.CrossRefGoogle ScholarPubMed
Sowers, T., Bender, M., Labeyrie, L., Martinson, D., Jouzel, J., Raynaud, D., Pichon, J.-J., and Korotkevich, Y. 135,000-year Vostok-SPECMAP common temporal framework. Paleoceanography 8, (1993). 737760.Google Scholar
Stewart, J.R. Intraspecific variation in modern and Quaternary European Lagopus . Smithsonian Contributions to Paleobiology 89, (2000). 5968.Google Scholar
Stiner, M.C. Honor among Thieves—A Zooarchaeological Study of Neandertal Sociology. (1994). Princeton Univ. Press, Princeton.Google Scholar
Stiner, M.C. Palaeolithic mollusc exploitation at Riparo Moche (Balzi Rossi, Italy); Food and ornaments from the Aurignacian through the Epigravettian. Antiquity 73, (1999). 735754.Google Scholar
Thompson, S.L., and Pollard, D. Greenland and Antarctic mass balances for present and doubled atmospheric CO2 from the GENESIS version-2 global climate model. Journal of Climatology 10, (1997). 87128900.Google Scholar
van Andel, T.H. Middle and Upper Palaeolithic environments and the calibration of 14C dates beyond 10,000 BP. Antiquity 72, (1998). 2633.Google Scholar
van Andel, T.H., and Tzedakis, P.C. Palaeolithic landscapes of Europe and environs, 150,000–25,000 years ago. Quaternary Science Reviews 15, (1996). 481500.Google Scholar
van Andel, T.H., and Tzedakis, P.C. Priority and opportunity: Reconstructing the European Middle Palaeolithic climate and landscape. Bayley, J. Science in Archaeology: An Agenda for the Future. (1998). English Heritage, London. 3746.Google Scholar
Völker, A.H., Sarnthein, M., Grootes, P., Erlenkeuser, H., Laj, C., Mazaud, A., Nadeau, M.-J., and Schleicher, X.X. Correlation of marine 14C ages from the Nordic Seas with the GISP2 isotope record: Implications for radiocarbon calibration beyond 25 ka BP. Radiocarbon 40, (1998). 517534.Google Scholar