Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-2xdlg Total loading time: 0 Render date: 2024-06-14T06:09:00.622Z Has data issue: false hasContentIssue false

4 - Hominins on the move: An assessment of anthropogenic shaping of environments in the Palaeolithic

from II - Origins: Species Movements in the Pleistocene

Published online by Cambridge University Press:  04 May 2017

Michael Petraglia
Affiliation:
Max Planck Institute
Nicole Boivin
Affiliation:
Max Planck Institute for the Science of Human History, Jena
Rémy Crassard
Affiliation:
Centre National de la Recherche Scientifique (CNRS), Lyon
Michael Petraglia
Affiliation:
Max Planck Institute for the Science of Human History, Jena
Get access

Summary

Abstract

Hominin dispersals in the Pliocene and Pleistocene have led to repeated range expansions of multiple human species, some with significant niche constructing behaviours. There is little doubt that humans have dramatically transformed global ecosystems since the adoption of agriculture by many societies in the Holocene. Beyond megafaunal extinctions, however, little attention has been paid to how pre-Holocene societies and our earlier hominin ancestors may have modified ecosystems as a consequence of subsistence-related activities and other pursuits. Evidence is reviewed here to demonstrate that the subsistence activities of hominins did in fact have an effect on local and regional environments as humans expanded their niches and territorial ranges in the Pliocene and Pleistocene. Evidence for the transformation of local ecologies in the Upper Palaeolithic of Europe is particularly convincing. Hominins also shaped their habitats through the use of fire and through the procurement and quarrying of raw materials for stone tool manufacture. Anthropogenic transformation of the natural world would appear to have begun in the Pliocene and Pleistocene, albeit on a different scale than in later periods.

Keywords: Human evolution, ecosystem, Palaeolithic, subsistence behaviour, landscape modification

INTRODUCTION

Humans today are shaping environments more than ever before in recorded history. The ways in which burgeoning human populations are contributing to the alteration of global climate and regional and local environments are the focus of investigation and interest by a diverse range of researchers, government organisations and policy makers. Most researchers have assumed that the transformation of Earth's climate and ecosystems by humans is largely a modern phenomenon. Yet, archaeologists and others are increasingly demonstrating that agricultural and pastoral societies significantly shaped environments over the course of the Holocene. While attention has thus been drawn to the ways in which human societies have altered global environments over the last 10,000 years, however, relatively little research has focused on studying how earlier human populations contributed to shaping ecosystems, beyond the intense debates surrounding the potential human contribution to the megafaunal extinctions that began ca. 50,000 years ago. Indeed, and perhaps somewhat surprisingly, Palaeolithic archaeologists have not yet made a concerted effort to examine the early anthropogenic transformation of ecosystems and physical landscapes, despite ethnographic observations suggesting that hunter-gatherers engaged in not insignificant environmental alteration.

Type
Chapter
Information
Human Dispersal and Species Movement
From Prehistory to the Present
, pp. 90 - 118
Publisher: Cambridge University Press
Print publication year: 2017

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

Alperson-Afil, N., Sharon, G., Kislev, M., Melamed, Y., Zohar, I., Ashkenazi, S., Rabinovich, R. et al. 2009. Spatial organization of hominin activities at Gesher Benot Ya'aqov, Israel. Science 326: 1677–1680.Google Scholar
Archer, W., Braun, D.R., Harris, J.W., McCoy, J.T., and Richmond, B.G. 2014. Early Pleistocene aquatic resource use in the Turkana Basin. Journal of Human Evolution 77: 74–87.Google Scholar
Archibald, S., Staver, A.C., and Levin, S.A. 2012. Evolution of human-driven fire regimes in Africa. Proceedings of the National Academy of Sciences 109: 847–852.Google Scholar
Atkinson, Q.D., Gray, R.D., and Drummond, A.J. 2008. MtDNA variation predicts population size in humans and reveals a major southern Asian chapter in human prehistory. Molecular Biology and Evolution 25: 468–474.Google Scholar
Bar-Yosef, O. 1994. The Lower Paleolithic of the Near East. Journal of World Prehistory 8: 211–265.Google Scholar
Ben-Dor, M., Gopher, A., Hershkovitz, I., and Barkai, R. 2011. Man the fat hunter: the demise of Homo erectus and the emergence of a new hominin lineage in the Middle Pleistocene (ca. 400 kyr) Levant. PLoS ONE: e28689.
Bentsen, S.E. 2014. Using pyrotechnology: fire-related features and activities with a focus on the African Middle Stone Age. Journal of Archaeological Research 22: 141–175.Google Scholar
Beyene, Y., Katoh, S., WoldeGabriel, G., Hart, W.K., Uto, K., Sudo, M., Hyodoi, M., et al. 2013. The characteristics and chronology of the earliest Acheulean at Konso, Ethiopia. Proceedings of the National Academy of Sciences 110: 1584–1591.Google Scholar
Bicho, N. and Haws, J. 2008. At the land's end: marine resources and the importance of fluctuations in the coastline in the prehistoric hunter–gatherer economy of Portugal. Quaternary Science Reviews 27: 2166–2175.Google Scholar
Bicho, N., Manne, T., Marreiros, J., Cascalheira, J., Pereira, T., Tátá, F., Évora, M., et al. 2013. The ecodynamics of the first modern humans in Southwestern Iberia: The case of Vale Boi, Portugal. Quaternary International 318: 102–116.Google Scholar
Bird, M.I., Hutley, L.B., Lawes, M.J., Lloyd, J.O.N., Luly, J.G., Ridd, P.V., Roberts, R.G., et al. 2013. Humans, megafauna and environmental change in tropical Australia. Journal of Quaternary Science 28: 439–452.Google Scholar
Blasco, R. and Peris, J.F. 2012. A uniquely broad spectrum diet during the Middle Pleistocene at Bolomor Cave (Valencia, Spain). Quaternary International 252: 16–31.Google Scholar
Blasco, R., Rosell, J., Smith, K.T., Maul, L.C., Sañudo, P., Barkai, R., and Gopher, A. 2016. Tortoises as a dietary supplement: a view from the Middle Pleistocene site of Qesem Cave, Israel. Quaternary Science Reviews 133: 165–182.Google Scholar
Bocherens, H., Bridault, A., Drucker, D.G., Hofreiter, M., Münzel, S.C., Stiller, M., and van Plicht, J. 2014. The last of its kind? Radiocarbon, ancient DNA and stable isotope evidence from a late cave bear (Ursus spelaeus ROSENMÜLLER, 1794) from Rochedane (France). Quaternary International 339: 179–188.Google Scholar
Braje, T.J. and Erlandson, J.M. 2013. Human acceleration of animal and plant extinctions: A Late Pleistocene, Holocene, and Anthropocene continuum. Anthropocene 4: 14–23.Google Scholar
Briggs, A.W., Good, J.M., Green, R.E., Krause, J., Maricic, T., Stenzel, U., Lalueza-Fox, C., et al. 2009. Targeted retrieval and analysis of five Neandertal mtDNA genomes. Science 325: 318–321.Google Scholar
Brown, K.S., Marean, C.W., Jacobs, Z., Schoville, B.J., Oestmo, S., Fisher, E.C., Bernatchez, J., et al. 2012. An early and enduring advanced technology originating 71,000 years ago in South Africa. Nature 491: 590–593.Google Scholar
Brunet, M., Beauvilain, A., Coppens, Y., Heintz, E., Moutaye, A.H.E., and Pilbeam, D. 1996. Australopithecus bahrelghazali, une nouvelle espèce d'Hominidé ancien de la région de Koro Toro (Tchad). Comptes rendus de l'Académie des Sciences, Série 2, Sciences de la Terre et des Planètes 322: 907–913.Google Scholar
Brunet, M., Guy, F., Pilbeam, D., Taisso Mackaye, H., Likius, A., Ahounta, D., Beauvilain, A., et al. 2002. A new hominid from the Upper Miocene of Chad, Central Africa. Nature 418: 145–151.Google Scholar
Burney, D.A. 1995. Historical perspectives on human‐assisted biological invasions. Evolutionary Anthropology 4: 216–221.Google Scholar
Carrión, J.S., Finlayson, C., Fernández, S., Finlayson, G., Allué, E., López-Sáez, J.A., López-García, P., et al. 2008. A coastal reservoir of biodiversity for Upper Pleistocene human populations: palaeoecological investigations in Gorham's Cave (Gibraltar) in the context of the Iberian Peninsula. Quaternary Science Reviews 27: 2118–2135.Google Scholar
Clark, J.L. 2011. The evolution of human culture during the later Pleistocene: using fauna to test models on the emergence and nature of ‘modern’ human behavior. Journal of Anthropological Archaeology 30: 273–291.Google Scholar
Conard, N.J., Serangeli, J., Böhner, U., Starkovich, B.M., Miller, C.E., Urban, B., and Van Kolfschoten, T. 2015. Excavations at Schöningen and paradigm shifts in human evolution. Journal of Human Evolution 89: 1–17.Google Scholar
Conard, N.J., Bolus, M., Goldberg, P., and Münzel, S.C. 2006. The last Neanderthals and first modern humans in the Swabian Jura. In When Neanderthals and Modern Humans Met, ed. Conard, N.J., pp. 305–341. Tübingen: Tübingen Publications in Prehistory.
Crutzen, P.J. and Stoermer, E.F. 2000. The Anthropocene. IGBP Newsletter 41: 12.Google Scholar
Davidson, A.M., Jennions, M., and Nicotra, A.B. 2011. Do invasive species show higher phenotypic plasticity than native species and, if so, is it adaptive? A meta-analysis. Ecology Letters 14: 419–431.Google Scholar
Dennell, R. 2003. Dispersal and colonisation, long and short chronologies: how continuous is the Early Pleistocene record for hominids outside East Africa? Journal of Human Evolution 45: 421–440.Google Scholar
Devès, M., Sturdy, D., Godet, N., King, G.C.P., and Bailey, G.N. 2014. Hominin reactions to herbivore distribution in the Lower Palaeolithic of the southern Levant. Quaternary Science Reviews 96: 140–160.Google Scholar
Drucker, D.G., Vercoutère, C., Chiotti, L., Nespoulet, R., Crépin, L., Conard, N.J., Münzel, S.C., et al. 2015. Tracking possible decline of woolly mammoth during the Gravettian In Dordogne (France) and the Ach Valley (Germany) using multi-isotope tracking (13C, 14C, 15N, 34S, 18O). Quaternary International 359–360:304–317.Google Scholar
Dusseldorp, G.L. 2012a. Studying prehistoric hunting proficiency: applying Optimal Foraging Theory to the Middle Palaeolithic and Middle Stone Age. Quaternary International 252: 3–15.Google Scholar
Dusseldorp, G.L. 2012b. Tracking the influence of technological change on Middle Stone Age hunting strategies in South Africa. Quaternary International 270: 70–79.Google Scholar
Edgeworth, M., deB Richter, D., Waters, C., Haff, P., Neal, C., and Price, S.J. 2015. Diachronous beginnings of the Anthropocene: the lower bounding surface of anthropogenic deposits. The Anthropocene Review 2: 33–58.
Espigares, M.P., Martínez-Navarro, B., Palmqvist, P., Ros-Montoya, S., Toro, I., Agustí, J., and Sala, R. 2013. Homo vs. Pachycrocuta: earliest evidence of competition for an elephant carcass between scavengers at Fuente Nueva-3 (Orce, Spain). Quaternary International 295: 113–125.Google Scholar
Fa, J.E., Stewart, J.R., Lloveras, L., and Vargas, J.M. 2013. Rabbits and hominin survival in Iberia. Journal of Human Evolution 64: 233–241.Google Scholar
Ferraro, J.V., Plummer, T.W., Pobiner, B.L., Oliver, J.S., Bishop, L.C., Braun, D.R., Ditchfield, P.W. et al. 2013. Earliest archaeological evidence of persistent hominin carnivory. PLoS ONE: e62174.
Ferring, R., Oms, O., Agustí, J., Berna, F., Nioradze, M., Shelia, T., Tappen, M., Vekua, E., et al. 2011. Earliest human occupations at Dmanisi (Georgian Caucasus) dated to 1.85–1.78 Ma. Proceedings of the National Academy of Sciences 108: 10432–10436.Google Scholar
Finlayson, C. 2009. The Humans Who Went Extinct. Oxford: Oxford University Press.
Flannery, K.V. 1969. Origins and ecological effects of early domestication in Iran and the Near East. In The Domestication and Exploitation of Plants and Animals, ed. Ucko, P.J. and Dimbleby, G.W., pp. 73–100. London: Duckworth.
Foley, R.A. 2013. Comparative evolutionary models and the ‘australopith radiations’. In The Paleobiology of Australopithecus, ed. Reed, K.E., Fleagle, J.G. and Leakey, R.E., pp. 163–174. Netherlands: Springer.
Foley, R.A., Maíllo-Fernández, J.M., and Lahr, M.M. 2013. The Middle Stone Age of the Central Sahara: biogeographical opportunities and technological strategies in later human evolution. Quaternary International 300: 153–170.Google Scholar
Foley, R.A. and Lahr, M.M. 2015. Lithic landscapes: early human impact from stone tool production on the Central Saharan environment. PLoS ONE 10(3): e0116482.Google Scholar
Foley, S.F., Gronenborn, D., Andreae, M.O., Kadereit, J.W., Esper, J., Scholz, D., Pöschl, U., et al. 2013. The Palaeoanthropocene – The beginnings of anthropogenic environmental change. Anthropocene 3: 83–88.Google Scholar
Fruth, B. and Hohmann, G. 1994. Nests: living artefacts of recent apes? Current Anthropology 35: 310–311.Google Scholar
Fuller, D.Q., van Etten, J., Manning, K., Castillo, C., Kingwell-Banham, E., Weisskopf, A., Qin, L., et al. 2011. The contribution of rice agriculture and livestock pastoralism to prehistoric methane levels: an archaeological assessment. The Holocene 21: 743–759.Google Scholar
Gaudzinski, S. 1996. On bovid assemblages and their consequences for the knowledge of subsistence patterns in the Middle Palaeolithic. Proceedings of the Prehistoric Society 62: 19–39.Google Scholar
Gaudzinski, S. and Roebroeks, W. 2000. Adults only. Reindeer hunting at the Middle Palaeolithic site Salzgitter Lebenstedt, northern Germany. Journal of Human Evolution 38: 497–521.Google Scholar
Glikson, A. 2013. Fire and human evolution: the deep-time blueprints of the Anthropocene. Anthropocene 3: 89–92.Google Scholar
Gopher, A. and Barkai, R. 2014. Middle Paleolithic open-air industrial areas in the Galilee, Israel: the challenging study of flint extraction and reduction complexes. Quaternary International 331: 95–102.Google Scholar
Goren-Inbar, N., Sharon, G., Melamed, Y., and Kislev, M. 2002. Nuts, nut cracking, and pitted stones at Gesher Benot Ya ‘aqov, Israel. Proceedings of the National Academy of Sciences 99: 2455–2460.Google Scholar
Gowlett, J.A. and Wrangham, R.W. 2013. Earliest fire in Africa: towards the convergence of archaeological evidence and the cooking hypothesis. Azania: Archaeological Research in Africa 48: 5–30.Google Scholar
Grayson, D.K. 2001. The archaeological record of human impacts on animal populations. Journal of World Prehistory 15: 1–68.Google Scholar
Grayson, D.K. 2007. Deciphering North American Pleistocene extinctions. Journal of Anthropological Research 63: 185–213.Google Scholar
Grayson, D.K. and Delpech, F. 2003. Ungulates and the middle-to-upper Paleolithic transition at Grotte XVI (Dordogne, France). Journal of Archaeological Science 30: 1633–1648.Google Scholar
Habeck-Fardy, A. and Nanson, G.C. 2014. Environmental character and history of the Lake Eyre Basin, one seventh of the Australian continent. Earth-Science Reviews 132: 39–66.Google Scholar
Hardy, B.L. and Moncel, M.H. 2011. Neanderthal use of fish, mammals, birds, starchy plants and wood 125–250,000 years ago. PLoS ONE: e23768.
Harmand, S., Lewis, J.E., Feibel, C.S., Lepre, C.J., Prat, S., Lenoble, A., Boës, X., et al. 2015. 3.3-million-year-old stone tools from Lomekwi 3, West Turkana, Kenya. Nature 521: 310–315.Google Scholar
Haws, J.A. 2012. Paleolithic socionatural relationships during MIS 3 and 2 in central Portugal. Quaternary International 264: 61–77.Google Scholar
Heldal, T. 2009. Constructing a quarry landscape from empirical data. General perspectives and a case study at the Aswan West Bank, Egypt. In ed. Abu-Jaber, N., Bloxam, E.G., Degryse, P. and Heldal, T., pp. 125–155. QuarryScapes: Ancient Stone Quarry Landscapes in the Eastern Mediterranean. Geological Survey of Norway Special Publication 12.
Henry, A.G., Brooks, A.S., and Piperno, D.R. 2014. Plant foods and the dietary ecology of Neanderthals and early modern humans. Journal of Human Evolution 69: 44–54.Google Scholar
Hockett, B. and Haws, J.A. 2005. Nutritional ecology and the human demography of Neandertal extinction. Quaternary International 137: 21–34.Google Scholar
Hockett, B.S. and Haws, J. 2009. Continuity in animal resource diversity in the Late Pleistocene human diet of Central Portugal. Before Farming 2(2): 1–14.Google Scholar
Hofreiter, M., Münzel, S., Conard, N.J., Pollack, J., Slatkin, M., Weiss, G., and Pääbo, S. 2007. Sudden replacement of cave bear mitochondrial DNA in the Late Pleistocene. Current Biology 17: R122–R123.Google Scholar
Hope, G. 2009. Environmental change and fire in the Owen Stanley ranges, Papua New Guinea. Quaternary Science Reviews 28: 2261–2276.Google Scholar
Huguet, R., Saladié, P., Cáceres, I., Díez, C., Rosell, J., Bennàsar, M., Blasco, R., et al. 2013. Successful subsistence strategies of the first humans in south-western Europe. Quaternary International 295: 168–182.Google Scholar
Hunt, C.O., Gilbertson, D.D., and Rushworth, G. 2012. A 50,000-year record of Late Pleistocene tropical vegetation and human impact in lowland Borneo. Quaternary Science Reviews 37: 61–80.Google Scholar
Iwase, A., Hashizume, J., Izuho, M., Takahashi, K., and Sato, H. 2012. Timing of megafaunal extinction in the late Late Pleistocene on the Japanese Archipelago. Quaternary International 255: 114–124.Google Scholar
Jennings, R.P., Shipton, C., Breeze, P., Cuthbertson, P., Bernal, M., Wedage, W.M.C.O., Drake, N.A., et al. 2015. Multi-scale Acheulean landscape survey in the Arabian Desert. Quaternary International 382: 58–81.Google Scholar
Jones, E.L. 2012. Upper Paleolithic rabbit exploitation and landscape patchiness: the Dordogne vs. Mediterranean Spain. Quaternary International 264: 52–60.Google Scholar
Karmin, M., Saag, L, Vicente, M., Wilson Sayres, M.A., Järve, M., Gerst Talas, U., Rootsi, S. et al. 2015. A recent bottleneck of Y chromosome diversity coincides with a global change in culture. Genome Research 25: 459–466.Google Scholar
Koch, P.L. and Barnosky, A.D. 2006. Late Quaternary extinctions: state of the debate. Annual Review of Ecology and Evolutionary Systems 37: 215–250.Google Scholar
Leakey, M. and Werdelin, L. 2010. Early Pleistocene mammals of Africa: background to dispersal. In Out of Africa I, ed. Fleagle, J.G., Shea, J.J., Grine, F.E., Baden, A.L. and Leakey, R.E., pp. 3–11. Netherlands: Springer.
Lewis, M.E. and Werdelin, L. 2007. Patterns of change in the Plio-Pleistocene carnivorans of eastern Africa. In Hominin Environments in the East African Pliocene: An Assessment of the Faunal Evidence, ed. Bobe, R., Alemseged, Z. and Behrensmeyer, A.K., pp. 77–105. Netherlands: Springer.
Lombard, M. and Phillipson, L. 2010. Indications of bow and stone-tipped arrow use 64000 years ago in KwaZulu-Natal, South Africa. Antiquity 84: 635–648.Google Scholar
Lorenzen, E.D., Nogués-Bravo, D., Orlando, L., Weinstock, J., Binladen, J., Marske, K.A., Ugan, A., et al. 2011. Species-specific responses of Late Quaternary megafauna to climate and humans. Nature 479: 359–364.Google Scholar
Martin, P.S. and Klein, R.G. (ed.) 1989. Quaternary Extinctions: A Prehistoric Revolution. The University of Arizona Press.
McCall, G.S. and Thomas, J.T. 2012. Still Bay and Howiesons Poort foraging strategies: Recent research and models of culture change. African Archaeological Review 29: 7–50.Google Scholar
McPherron, S.P., Alemseged, Z., Marean, C.W., Wynn, J.G., Reed, D., Geraads, D., Bobe, R., and Béarat, H.A. 2010. Evidence for stone-tool-assisted consumption of animal tissues before 3.39 million years ago at Dikika, Ethiopia. Nature 466: 857–860.Google Scholar
Mellars, P. 1996. The Neanderthal Legacy: An Archaeological Perspective from Western Europe. Princeton: Princeton University Press.
Mellars, P. 2011. The earliest modern humans in Europe. Nature 479: 483–485.Google Scholar
Mellars, P. and French, J.C. 2011. Tenfold population increase in western Europe at the Neandertal-to-modern human transition. Science 333: 623–627.Google Scholar
Miller, G.H., Fogel, M.L., Magee, J.W., Gagan, M.K., Clarke, S.J., and Johnson, B.J. 2005. Ecosystem collapse in Pleistocene Australia and a human role in megafaunal extinction. Science 309: 287–290.Google Scholar
Miller, G.H., Magee, J.W., Fogel, M.L., and Gagan, M.K. 2007. Detecting human impacts on the flora, fauna, and summer monsoon of Pleistocene Australia. Climate of the Past 3: 463–473.Google Scholar
Münzel, S.C. and Conard, N.J. 2004. Cave bear hunting in Hohle Fels cave in the Ach valley of the Swabian Jura. Revue de Paléobiologie 23: 877–885.Google Scholar
Münzel, S.C., Stiller, M., Hofreiter, M., Mittnik, A., Conard, N.J., and Bocherens, H. 2011. Pleistocene bears in the Swabian Jura (Germany): genetic replacement, ecological displacement, extinctions and survival. Quaternary International 245: 225–237.Google Scholar
Nikolskiy, P. and Pitulko, V. 2013. Evidence from the Yana Palaeolithic site, Arctic Siberia, yields clues to the riddle of mammoth hunting. Journal of Archaeological Science 40: 4189–4197.Google Scholar
Nogués-Bravo, D., Ohlemüller, R., Batra, P., and Araújo, M.B. 2010. Climate predictors of late Quaternary extinctions. Evolution 64: 2442–2449.Google Scholar
Norton, C.J., Kondo, Y., Ono, A., Zhang, Y., and Diab, M.C. 2010. The nature of megafaunal extinctions during the MIS 3–2 transition in Japan. Quaternary International 211: 113–122.Google Scholar
O'Connell, J.F. and Allen, J. 2015. The process, biotic impact, and global implications of the human colonization of Sahul about 47,000 years ago. Journal of Archaeological Science 56: 73–84.Google Scholar
O'Regan, H. J., Turner, A., and Wilkinson, D.M. 2002. European Quaternary refugia: a factor in large carnivore extinction? Journal of Quaternary Science 17: 789–795.Google Scholar
Panger, M.A., Brooks, A.S., Richmond, B.G., and Wood, B. 2002. Older than the Oldowan? Rethinking the emergence of hominin tool use. Evolutionary Anthropology 11: 235–245.Google Scholar
Pante, M.C. 2013. The larger mammal fossil assemblage from JK2, Bed III, Olduvai Gorge, Tanzania: implications for the feeding behavior of Homo erectus . Journal of Human Evolution 64: 68–82.Google Scholar
Petraglia, M., Clarkson, C., Boivin, N., Haslam, M., Korisettar, R., Chaubey, G., Ditchfield, P., et al. 2009. Population increase and environmental deterioration correspond with microlithic innovations in South Asia ca. 35,000 years ago. Proceedings of the National Academy of Sciences 106: 12261–12266.Google Scholar
Petraglia, M., LaPorta, P., and Paddayya, K. 1999. The first Acheulian quarry in India: stone tool manufacture, biface morphology, and behaviors. Journal of Anthropological Research 55: 39–70.Google Scholar
Plummer, T. 2004. Flaked stones and old bones: biological and cultural evolution at the dawn of technology. American Journal of Physical Anthropology 125(S39): 118–164.Google Scholar
Potts, R. 2013. Hominin evolution in settings of strong environmental variability. Quaternary Science Reviews 73: 1–13.Google Scholar
Prescott, G.W., Williams, D.R., Balmford, A., Green, R.E., and Manica, A. 2012. Quantitative global analysis of the role of climate and people in explaining late Quaternary megafaunal extinctions. Proceedings of the National Academy of Sciences 109: 4527–4531.Google Scholar
Rabinovich, R., Gaudzinski-Windheuser, S., Kindler, L., and Goren-Inbar, N. 2012. The Acheulean Site of Gesher Benot Ya'aqov Volume III: Mammalian Taphonomy, the Assemblages of Layers V-5 and V-6. Dordrecht: Springer.
Rhodin, A.G.J., Thomson, S., Georgalis, G.L., Karl, H.-V., Danilov, I.G., Takahashi, A., Fuente, M.S., et al. 2015. Turtles and tortoises of the world during the rise and global spread of humanity: first checklist and review of extinct Pleistocene and Holocene Chelonians. In A Compilation Project of the IUCN/SSC Tortoise and Freshwater Turtle Specialist Group, ed. Rhodin, A.G.J., Pritchard, P.C.H., Dijk, P.P. van, Saumure, R.A., Buhlmann, K.A., Iverson, J.B. and Mittermeier, R.A., pp. 000e.1–000e.66. Chelonian Research Monographs, Chelonian Research Foundation.
Ricciardi, A. 2007. Are modern biological invasions an unprecedented form of global change? Conservation Biology 21: 329–336.Google Scholar
Roberts, M.B. and Parfitt, S.A. 1999. Boxgrove: A Middle Pleistocene Hominid Site at Eartham Quarry, Boxgrove. English Heritage.
Roberts, P., Delson, E., Miracle, P., Ditchfield, P., Roberts, R.G., Jacobs, Z., Blinkhorn, J., et al. 2014. Continuity of mammalian fauna over the last 200,000 y in the Indian subcontinent. Proceedings of the National Academy of Sciences 111: 5848–5853.Google Scholar
Roebroeks, W. and Villa, P. 2011. On the earliest evidence for habitual use of fire in Europe. Proceedings of the National Academy of Sciences 108: 5209–5214.Google Scholar
Ruddiman, W.F. 2003. The anthropogenic greenhouse era began thousands of years ago. Climatic Change 61: 261–293.Google Scholar
Ruddiman, W.F., Ellis, E.C., Kaplan, J.O., and Fuller, D.Q. 2015. Defining the epoch we live in. Science 348: 38–39.Google Scholar
Rule, S., Brook, B.W., Haberle, S.G., Turney, C.S., Kershaw, A.P., and Johnson, C.N. 2012. The aftermath of megafaunal extinction: ecosystem transformation in Pleistocene Australia. Science 335: 1483–1486.Google Scholar
Salazar-García, D.C., Power, R.C., Sanchis Serra, A., Villaverde, V., Walker, M.J., and Henry, A.G. 2013. Neanderthal diets in central and southeastern Mediterranean Iberia. Quaternary International 318: 3–18.Google Scholar
Saltré, F., Rodríguez-Rey, M., Brook, B.W., Johnson, C.N., Turney, C.S., Alroy, J., Cooper, A., et al. 2016. Climate change not to blame for late Quaternary megafauna extinctions in Australia. Nature Communications 7: 10511.Google Scholar
Sampson, G. 2001. An Acheulian settlement pattern in the Upper Karoo region of South Africa. In A Very Remote Period Indeed: Papers on the Palaeolithic Presented to Derek Roe, ed. Milliken, S. and Cook, J., pp. 28–36. Oxford: Oxbow Books.
Sampson, G. 2006. Acheulean quarries at hornfels outcrops in the Upper Karoo region of South Africa. In Axe Age, Acheulian Toolmaking – from Quarry to Discard, ed. Goren-Inbar, N. and Sharon, G., pp. 75–107. London: Equinox.
Sandom, C., Faurby, S., Sandel, B., and Svenning, J.-C. 2014. Global late Quaternary megafauna extinctions linked to humans, not climate change. Proceedings of the Royal Society B 281: 20133254.Google Scholar
Semaw, S., Renne, P., Harris, J.W.K., Feibel, C.S., Bernor, R.L., Fesseha, N., and Mowbray, K. 1997. 2.5-million-year-old stone tools from Gona, Ethiopia. Nature 385: 333–336.Google Scholar
Senut, B., Pickford, M., Gommery, D., Mein, P., Cheboi, K., and Coppens, Y. 2001. First hominid from the Miocene (Lukeino formation, Kenya). Comptes Rendus de l'Académie des Sciences-Series IIA-Earth and Planetary Science 332(2): 137–144.Google Scholar
Shea, J.J. 2006. The origins of lithic projectile point technology: evidence from Africa, the Levant, and Europe. Journal of Archaeological Science 33: 823–846.Google Scholar
Shea, J.J. and Sisk, M.L. 2010. Complex projectile technology and Homo sapiens dispersal into western Eurasia . PaleoAnthropology 2010 100–122.
Shipman, P. 2015. How do you kill 86 mammoths? Taphonomic investigations of mammoth megasites. Quaternary International 359–360:38–46.Google Scholar
Smith, B.D. and Zeder, M.A. 2013. The onset of the Anthropocene. Anthropocene 4: 8–13.Google Scholar
Stewart, F.A., Piel, A.K., and McGrew, W.C. 2011. Living archaeology: artefacts of specific nest site fidelity in wild chimpanzees. Journal of Human Evolution 61: 388–395.Google Scholar
Stewart, J.R. 2004. Neanderthal–modern human competition? A comparison between the mammals associated with Middle and Upper Palaeolithic industries in Europe during OIS 3. International Journal of Osteoarchaeology 14: 178–189.Google Scholar
Stiller, M., Baryshnikov, G., Bocherens, H., Grandal d'Anglade, A., Hilpert, B., Münzel, S.C., Pinhasi, R., et al. 2010. Withering away – 25,000 years of genetic decline preceded cave bear extinction. Molecular Biology and Evolution 27: 975–978.Google Scholar
Stimpson, C. 2010. Late Quaternary environments and human impact in northern Borneo: the evidence from the bird (Aves) and Bat (Mammalia: Chiroptera) faunas from the archaeology of the Great Cave of Niah, Sarawak. Unpublished PhD thesis, University of Cambridge, Cambridge, UK.
Stiner, M.C. 2001. Thirty years on the ‘Broad Spectrum Revolution’ and paleolithic demography. Proceedings of the National Academy of Sciences 98: 6993–6996.Google Scholar
Stiner, M.C., Munro, N.D., and Surovell, T.A. 2000. The tortoise and the hare. Current Anthropology 41: 39–79.Google Scholar
Strait, D.S. and Wood, B.A. 1999. Early hominid biogeography. Proceedings of the National Academy of Sciences 96: 9196–9200.Google Scholar
Straus, L.G. 1982. Carnivores and cave sites in Cantabrian Spain. Journal of Anthropological Research 38: 75–96.Google Scholar
Stringer, C.B., Finlayson, J.C., Barton, R.N.E., Fernández-Jalvo, Y., Cáceres, I., Sabin, R.C., Rhodes, E.J., et al. 2008. Neanderthal exploitation of marine mammals in Gibraltar. Proceedings of the National Academy of Sciences 105: 14319–14324.Google Scholar
Summerhayes, G.R., Leavesley, M., Fairbairn, A., Mandui, H., Field, J., Ford, A., and Fullagar, R. 2010. Human adaptation and plant use in highland New Guinea 49,000 to 44,000 years ago. Science 330: 78–81.Google Scholar
Surovell, T., Waguespack, N., and Brantingham, J. 2005. Global archaeological evidence for proboscidean overkill. Proceedings of the National Academy of Sciences 102: 6231–6236.Google Scholar
Thevenon, F., Bard, E., Williamson, D., and Beaufort, L. 2004. A biomass burning record from the West Equatorial Pacific over the last 360 ky: methodological, climatic and anthropic implications. Palaeogeography, Palaeoclimatology, Palaeoecology 213: 83–99.Google Scholar
Thieme, H. 1997. Lower Palaeolithic hunting spears from Germany. Nature 385: 807–810.Google Scholar
Turney, C.S., Kershaw, A.P., Moss, P., Bird, M.I., Fifield, L.K., Cresswell, R.G., Santos, G.M., et al. 2001. Redating the onset of burning at Lynch's Crater (North Queensland): implications for human settlement in Australia. Journal of Quaternary Science 16: 767–771.Google Scholar
Van Schaik, C.P., Ancrenaz, M., Borgen, G., Galdikas, B., Knott, C.D., Singleton, I., Suzuki, A., et al. 2003. Orangutan cultures and the evolution of material culture. Science 299: 102–105.Google Scholar
Vermeersch, P.M. (ed.) 2002. Palaeolithic Quarrying Sites in Upper and Middle Egypt. No. 4., Leuven University Press.
Vermeij, G.J. 1991. When biotas meet: understanding biotic interchange. Science 253: 1099–1104.Google Scholar
Vermeij, G.J. 2005. Invasion as expectation. In Species Invasions: Insights into Ecology, Evolution, and Biogeography, ed. Sax, D.F., Stachowicz, J.J. and Gaines, S.D., pp. 315–339. Sunderland, Massachusetts: Sinauer Associates.
Villaverde Bonilla, V., Martínez-Valle, R., Guillem-Calatayud, P.M., Badal, E., Zalbidea, L., and García, R. 1997. Mobility and the role of small game in the Middle Paleolithic of the central region of the Spanish Mediterranean: a comparison of Cova Negra with other Paleolithic deposits. In The Last Neandertals, The First Anatomically Modern Humans: A Tale about Human Diversity, Cultural Change and Human Evolution, ed. Carbonell, E. and Vaquero, M., pp. 267–288. Tarragona: Spain.
Wadley, L. 2010. Were snares and traps used in the Middle Stone Age and does it matter? A review and a case study from Sibudu, South Africa. Journal of Human Evolution 58: 179–192.Google Scholar
Wells, J.C. and Stock, J.T. 2007. The biology of the colonizing ape. American Journal of Physical Anthropology 134: 191–222.Google Scholar
Williams, M., Zalasiewicz, J., Haff, P.K., Schwägerl, C., Barnosky, A.D., and Ellis, E.C. 2015. The Anthropocene biosphere. The Anthropocene Review 2: 196–219.Google Scholar
Zalasiewicz, J., Williams, M., Steffen, W., and Crutzen, P. 2010. The new world of the Anthropocene. Environmental Science & Technology 44: 2228–2231.Google Scholar
Zeder, M.A. 2012. The broad spectrum revolution at 40: resource diversity, intensification, and an alternative to optimal foraging explanations. Journal of Anthropological Archaeology 31: 241–264.Google Scholar

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.

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.

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.

Available formats
×