Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-r6qrq Total loading time: 0 Render date: 2024-04-28T16:29:51.968Z Has data issue: false hasContentIssue false

Bibliography

Published online by Cambridge University Press:  16 March 2023

John J. Shea
Affiliation:
State University of New York, Stony Brook
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
The Unstoppable Human Species
The Emergence of Homo Sapiens in Prehistory
, pp. 320 - 336
Publisher: Cambridge University Press
Print publication year: 2023

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

Adams, J. L. 2014a. Ground Stone Analysis: An Anthropological Approach. Salt Lake City: University of Utah Press.Google Scholar
Adams, S. 2014b. How to Fail at Almost Everything and Still Win Big: Kind of the Story of My Life. New York: Portfolio (Penguin/Random House).Google Scholar
Aiello, L. C., and Wheeler, P.. 1995. The expensive-tissue hypothesis: The brain and digestive system in human and primate evolution. Current Anthropology 36:199221.CrossRefGoogle Scholar
Alton, J., and Alton, A.. 2013. The Survival Medicine Handbook: A Guide for When Help Is Not on the Way, 2nd ed. Lexington, KY: Doom and Bloom.Google Scholar
Ambrose, S. H. 1998. Late Pleistocene human population bottlenecks, volcanic winter, and the differentiation of modern humans. Journal of Human Evolution 34:623651.Google Scholar
Ambrose, S. H. 2002. “Small Things Remembered: Origins of Early Microlithic Industries in Sub-Saharan Africa,” in Thinking Small: Global Perspectives on Microlithization. Edited by Elston, R. G. and Kuhn, S. L., pp. 930. Washington, DC: American Anthropological Association (Archaeological Paper No. 12).Google Scholar
Anderson-Gerfaud, P. C. 1990. “Aspects of Behaviour in the Middle Palaeolithic: Functional Analysis of Stone Tools from Southwest France,” in The Emergence of Modern Humans. Edited by Mellars, P. A., pp. 389418. Edinburgh: Edinburgh University Press.Google Scholar
Andrefsky, W. J. 2005. Lithics: Macroscopic Approaches to Analysis, 2nd ed. New York: Cambridge University Press.Google Scholar
Aranguren, B., Revedin, A., et al. 2018. Wooden tools and fire technology in the early Neanderthal site of Poggetti Vecchi (Italy). Proceedings of the National Academy of Sciences 115:2054.Google Scholar
Arens, W. 1979. The Man-Eating Myth: Anthropology and Anthropophagy. New York: Oxford University Press.Google Scholar
Ashley, C. Z., and Grillo, K. M.. 2015. Archaeological ceramics from eastern Africa: Past approaches and future directions. Azania: Archaeological Research in Africa 50:460480.Google Scholar
Aubert, M., Lebe, R., et al. 2019. Earliest hunting scene in prehistoric art. Nature 576:442445.Google Scholar
Aubert, M., Setiawan, P., et al. 2018. Palaeolithic cave art in Borneo. Nature 564:254257.Google Scholar
Backwell, L., Bradfield, J., et al. 2018. The antiquity of bow-and-arrow technology: Evidence from Middle Stone Age layers at Sibudu Cave. Antiquity 92:289303.CrossRefGoogle Scholar
Bae, C. J., Douka, K., et al. 2017. On the origin of modern humans: Asian perspectives. Science 358:eaai9067.Google Scholar
Bahn, P. G. 2016. Images of the Ice Age. New York: Oxford University Press.Google Scholar
Bailey, R., and Peacock, N.. 1988. “Efe Pygmies of Northeast Zaire: Subsistence Strategies in the Ituri Forest,” in Coping with Uncertainty in the Food Supply. Edited by de Garine, I. and Harrison, G., pp. 88117. Oxford: Oxford University Press.Google Scholar
Bar-Yosef, O. 2002. The Upper Paleolithic Revolution. Annual Review of Anthropology 31:363393.Google Scholar
Bar-Yosef, O., and Belfer-Cohen, A.. 2001. From Africa to Eurasia – Early dispersals. Quaternary International 75:1928.Google Scholar
Bar-Yosef, O., and Callander, J.. 1999. The woman from Tabun: Garrod’s doubts in historical perspective. Journal of Human Evolution 37:879885.Google Scholar
Bar-Yosef, O., Eren, M., et al. 2012. Were bamboo tools made in prehistoric Southeast Asia? An experimental view from South China. Quaternary International 269:912.CrossRefGoogle Scholar
Bar-Yosef Mayer, D. 2005. The exploitation of shells as beads in the Palaeolithic and Neolithic of the Levant. Paléorient 31:176185.Google Scholar
Bar-Yosef Mayer, D. E., Groman-Yaroslavski, I., Bar-Yosef, O., Hershkovitz, I., Kampen-Hasday, A., Vandermeersch, B., Zaidner, Y., and Weinstein-Evron, M.. 2020. On holes and strings: Earliest displays of human adornment in the Middle Palaeolithic. PLoS ONE 15(7):e0234924.Google Scholar
Barham, L. 2002. Backed tools in Middle Pleistocene Central Africa and their evolutionary significance. Journal of Human Evolution 43:585603.Google Scholar
Barham, L. 2013. From Hand to Handle: The First Industrial Revolution. New York: Oxford University Press.Google Scholar
Barham, L., and Mitchell, P.. 2008. The First Africans: African Archaeology from the Earliest Toolmakers to Most Recent Foragers. New York: Cambridge University Press.Google Scholar
Barham, L., Simms, M. J., et al. 2000. “Twin Rivers, excavation and behavioural record,” in The Middle Stone Age of Zambia, South Central Africa. Edited by Barham, L., pp. 165216. Bristol: Western Academic & Specialist Press.Google Scholar
Barker, G. 2014. “Niah Caves: Role in Human Evolution,” in Encyclopedia of Global Archaeology. Edited by Smith, C., pp. 52825289. New York: Springer New York.Google Scholar
Barton, C. M., and Clark, G. A.. 2021. From artifacts to cultures: Technology, society, and knowledge in the Upper Paleolithic. Journal of Paleolithic Archaeology 4 (2):121.Google Scholar
Becerra-Valdivia, L., and Higham, T.. 2020. The timing and effect of the earliest human arrivals in North America. Nature 584:9397.CrossRefGoogle ScholarPubMed
Bednarik, R. G. 2001. Replicating the first known sea travel by humans: The Lower Pleistocene crossing of Lombok Strait. Human Evolution 16:229242.Google Scholar
Bednarik, R. G. 2003. Seafaring in the Pleistocene. Cambridge Archaeological Journal 13:4166.CrossRefGoogle Scholar
Begley, C. 2021. The Next Apocalypse: The Art and Science of Survival. New York: Basic Books.Google Scholar
Bellwood, P. 1987. The Prehistory of Island Southeast Asia: A Multidisciplinary Review of Recent Research. Journal of World Prehistory 1:171224.Google Scholar
Bellwood, P. 2013. First Migrants: Ancient Migration in Global Perspective. New York: Wiley Blackwell.Google Scholar
Ben Arous, E., Philippe, A., et al. 2022. An improved chronology for the Middle Stone Age at El Mnasra cave, Morocco. PLoS ONE 17:e0261282.Google Scholar
Ben-Dor, M., and Barkai, R.. 2020. The importance of large prey animals during the Pleistocene and the implications of their extinction on the use of dietary ethnographic analogies. Journal of Anthropological Archaeology 59:101192.Google Scholar
Bendictow, O. J. 2021. The Complete History of the Black Death. Woodbridge: Boydell Press.Google Scholar
Bennett, M. R., Bustos, D., et al. 2021. Evidence of humans in North America during the Last Glacial Maximum. Science (American Association for the Advancement of Science) 373:15281531.Google Scholar
Berger, T. D., and Trinkaus, E.. 1995. Patterns of trauma among the Neandertals. Journal of Archaeological Science 22:841852.CrossRefGoogle Scholar
Bergman, C., and Newcomer, M. H.. 1983. Flint arrowhead breakage: Examples from Ksar Akil, Lebanon. Journal of Field Archaeology 10:238243.Google Scholar
Bergström, A., Stringer, C., et al. 2021. Origins of modern human ancestry. Nature 590:229237.Google Scholar
Beyin, A. 2006. The Bab al Mandab vs the Nile-Levant: An appraisal of the two dispersal routes for early modern humans out of Africa. African Archaeological Review 23:530.Google Scholar
Béyries, S. 1987. Variabilité de l’industrie lithique au Moustérien: Approche fonctionelle sur quelques gisements français. British Archaeological Reports International Series, no. 328. Oxford: Oxbow Books.CrossRefGoogle Scholar
Binford, L. R. 1962. Archaeology as anthropology. American Antiquity 28:217225.Google Scholar
Binford, L. R. 1979. Organization and formation processes: Looking at curated technologies. Journal of Anthropological Research 35:255273.Google Scholar
Binford, L. R. 1981. Bones: Ancient Men and Modern Myths. New York: Academic Press.Google Scholar
Bingham, P. M., and Souza, J.. 2013. Theory testing in Prehistoric North America: Fruits of one of the world’s great archeological natural laboratories. Evolutionary Anthropology: Issues, News, and Reviews 22:145153.CrossRefGoogle ScholarPubMed
Bingham, P. M., Souza, J., et al. 2013. Introduction: Social complexity and the bow in the prehistoric North American record. Evolutionary Anthropology: Issues, News, and Reviews 22:8188.CrossRefGoogle ScholarPubMed
Blinkhorn, J., Lucy, T., et al. 2022. Evaluating refugia in recent human evolution in Africa. Philosophical Transactions of the Royal Society B: Biological Sciences 377:20200485.Google Scholar
Blinkhorn, J., Zanolli, C., et al. 2021. Nubian Levallois technology associated with southernmost Neanderthals. Scientific Reports 11:2869.Google Scholar
Boaretto, E., Wu, X., et al. 2009. Radiocarbon dating of charcoal and bone collagen associated with early pottery at Yuchanyan Cave, Hunan Province, China. Proceedings of the National Academy of Sciences 106:9595.Google Scholar
Boas, F. 1940. Race, Language, and Culture. New York: Macmillan.Google Scholar
Boëda, E., Connan, J., et al. 1996. Bitumen as a hafting material on Middle Palaeolithic artefacts. Nature 380:336338.CrossRefGoogle Scholar
Boldurian, A. T. 2007. Clovis beveled rod manufacture: An elephant bone experiment. North American Archaeologist 28:2957.CrossRefGoogle Scholar
Bordes, F. 1968. The Old Stone Age. New York: McGraw-Hill.Google Scholar
Bostrum, N. 2014. Superintelligence: Paths, Dangers, Strangers. New York: Oxford University Press.Google Scholar
Bostrum, N., and Cirkovic, M.. Editors. 2011. Global Catastrophic Risks. New York: Oxford University Press.Google Scholar
Bowles, S. 2009. Did warfare among ancestral hunter-gatherers affect the evolution of human social behaviors? Science 324:12931298.Google Scholar
Bradford, W., and Morrison, S. E. e. 1651 (1952). Of Plymouth Plantation, 1620–1647. New York: Afred A. Knopf.Google Scholar
Brain, C. K. 1981. The Hunters or the Hunted? An Introduction to African Cave Taphonomy. Chicago, IL: University of Chicago Press.Google Scholar
Bramble, D. M., and Lieberman, D. E.. 2004. Endurance running and the evolution of Homo. Nature 432:354352.Google Scholar
Brandt, S. A., Fisher, E. C., et al. 2012. Early MIS 3 occupation of Mochena Borago Rockshelter, Southwest Ethiopian Highlands: Implications for Late Pleistocene archaeology, paleoenvironments and modern human dispersals. Quaternary International 274:3854.Google Scholar
Brannen, P. 2017. The Ends of the World: Volcanic Apocalypses, Lethal Oceans, and Our Quest to Understand Earth’s Past Mass Extinctions. New York: HarperCollins.Google Scholar
Brooks, A. S., Yellen, J. E., et al. 2018. Long-distance stone transport and pigment use in the earliest Middle Stone Age. Science 360:90.CrossRefGoogle ScholarPubMed
Brown, D. J. 2015. The Indifferent Stars Above: The Harrowing Saga of a Donner Party Bride. New York: William Morrow.Google Scholar
Brown, K. S., Marean, C. W., et al. 2009. Fire as an engineering tool of early modern humans. Science 325:859862.Google Scholar
Brumm, A. 2010. The Movius Line and the Bamboo Hypothesis: Early hominin stone technology in Southeast Asia. Lithic Technology 35:724.Google Scholar
Burdukiewicz, J. M., and Ronen, A.. Editors. 2003. Lower Paleolithic Small Tools in Europe and the Levant. British Archaeological Reports International Series, no. 1115. Oxford: Oxbow Books.CrossRefGoogle Scholar
Burroughs, W. J. 2005. Climate Change in Prehistory: The End of the Reign of Chaos. Cambridge: Cambridge University Press.Google Scholar
Callahan, E. 1979. The Basics of Biface Knapping in the Eastern Fluted Point Tradition. Archaeology of Eastern North America 7:1180.Google Scholar
Carrier, D. R. 1984. The energetic paradox of human running and hominid evolution. Current Anthropology 25(4):483495.Google Scholar
Cattelain, P. 1997. “Hunting during the Upper Paleolithic: Bow, spearthrower, or both?,” in Projectile Technology. Edited by Knecht, H., pp. 213240. New York: Plenum.CrossRefGoogle Scholar
Chase, P. G. 1989. “How Different was Middle Paleolithic Subsistence? A Zooarchaeological Perspective on the Middle to Upper Palaeolithic Transition,” in The Human Revolution: Behavioural and Biological Perspectives on the Origins of Modern Humans. Edited by Mellars, P. A. and Stringer, C. B., pp. 321337. Edinburgh: Edinburgh University Press.Google Scholar
Cheng, H., Zhang, H., et al. 2020. Timing and structure of the Younger Dryas event and its underlying climate dynamics. Proceedings of the National Academy of Sciences 117:23408.CrossRefGoogle ScholarPubMed
Cheynier, A. 1958. Impromptu sur la séquence des pointes du Paléolithique supérieur. Bulletin de la Société préhistorique de France 55:190205.Google Scholar
Chomsky, N. 2012. The Science of Language. New York: Cambridge University Press.Google Scholar
Christel, M. 1994. Catarrhine primates grasping small objects: Techniques and hand preferences. Current Primatology 3:3749.Google Scholar
Churchill, S. E. 1993. “Weapon Technology, Prey Size Selection, and Hunting Methods in Modern Hunter-Gatherers: Implications for Hunting in the Palaeolithic and Mesolithic,” in Hunting and Animal Exploitation in the Later Paleolithic and Mesolithic of Eurasia. Edited by Peterkin, G. L., Bricker, H., and Mellars, P. A., pp. 1124. Washington, DC: American Anthropological Association (Archaeological Paper No. 4).Google Scholar
Churchill, S. E. 2014. Thin on the Ground: Neandertal Biology, Archaeology and Ecology. New York: Wiley-Blackwell.Google Scholar
Churchill, S. E., and Smith, F. H.. 2000. Makers of the Early Aurignacian of Europe. Yearbook of Physical Anthropology 43:61115.Google Scholar
Clark, G. 1970. Aspects of Prehistory. Berkeley: University of California Press.Google Scholar
Clark, G. A. 1994. Migration as an explanatory concept in Paleolithic archaeology. Journal of Archaeological Method and Theory 1:305343.CrossRefGoogle Scholar
Clark, G. A., and Riel-Salvatore, J.. 2006. “Observations on Systematics in Paleolithic Archaeology,” in Transitions before the Transition. Edited by Hovers, E. and Kuhn, S. L., pp. 2957. New York: Plenum/Kluwer.Google Scholar
Clark, J. D. 1967. Atlas of African Prehistory. Chicago, IL: University of Chicago Press.Google Scholar
Clark, J. D. 1988. The Middle Stone Age of East Africa and the beginnings of regional identity. Journal of World Prehistory 2:235305.Google Scholar
Clark, J. D. Editor. 2001. Kalambo Falls Prehistoric Site, vol. III: The Earlier Cultures: Middle and Earlier Stone Age. Cambridge: Cambridge University Press.Google Scholar
Clarkson, C., Harris, C., et al. 2020. Human occupation of northern India spans the Toba super-eruption ~74,000 years ago. Nature Communications 11:961.CrossRefGoogle Scholar
Clarkson, C., Jacobs, Z., et al. 2017. Human occupation of northern Australia by 65,000 years ago. Nature 547:306.Google Scholar
Conard, N. J. 2006. “The Last Neanderthals and First Modern Humans in the Swabian Jura,” in When Neanderthals and Modern Humans Met. Edited by Conard, N. J., pp. 305341. Tübingen: Kerns Verlag.Google Scholar
Coqueugniot, H., Dutour, O., et al. 2014. Earliest cranio-encephalic trauma from the Levantine Middle Palaeolithic: 3D reappraisal of the Qafzeh 11 skull, consequences of pediatric brain damage on individual life condition and social care. PLoS ONE 9:e102822.Google Scholar
Crabtree, D. E. 1966. A stoneworker’s approach to analyzing and replicating the Lindenmeier Folsom. Tebiwa 9:339.Google Scholar
Crosby, A. W. 1986. Ecological Imperialism: The Biological Expansion of Europe, 900–1900. Cambridge: Cambridge University Press.Google Scholar
Cummings, V., Jordan, P., et al. 2014. The Oxford Handbook of the Archaeology and Anthropology of Hunter-Gatherers. New York: Oxford University Press.Google Scholar
Daniel, G. E., and Renfrew, C.. 1988. The Idea of Prehistory. Edinburgh: Edinburgh University Press.Google Scholar
Darwin, C. R. 1859. The Origin of Species. London: John Murray.Google Scholar
Dawkins, R. 1976. The Selfish Gene. London: Oxford University Press.Google Scholar
de Filippo, C., Bostoen, K., et al. 2012. Bringing together linguistic and genetic evidence to test the Bantu expansion. Proceedings of the Royal Society B: Biological Sciences 279:32563263.Google Scholar
de Mortillet, G. 1877. L’art dans les temps géologiques. Revue scientifique (second series):888892.Google Scholar
Defleur, A., White, T., et al. 1999. Neanderthal cannibalism at Moula-Guercy, Ardèche, France. Science 286:128131.CrossRefGoogle ScholarPubMed
Dent, R. J. 2007. “Seed Collecting and Fishing at the Shawnee Minisink Paleoindian Site Everyday Life in the Late Pleistocene,” in Foragers of the Terminal Pleistocene in North America. Edited by Walker, R. B and Driskell, B. N, pp. 116131. Lincoln: University of Nebraska Press.Google Scholar
d’Errico, F., Henshilwood, C., et al. 2003. Archaeological evidence for the emergence of language, symbolism, and music – An alternative multidisciplinary approach. Journal of World Prehistory 17:170.CrossRefGoogle Scholar
d’Errico, F., Henshilwood, C., et al. 2005. Nassarius kraussianus shell beads from Blombos Cave: Evidence for symbolic behaviour in the Middle Stone Age. Journal of Human Evolution 48:324.CrossRefGoogle ScholarPubMed
Dewar, R. E., Radimilahy, C., et al. 2013. Stone tools and foraging in northern Madagascar challenge Holocene extinction models. Proceedings of the National Academy of Sciences of the United States of America 110:1258312588.Google Scholar
Dibble, H. L. 1995. Middle Paleolithic scraper reduction: Background, clarification, and review of the evidence to date. Journal of Archaeological Method and Theory 2:299368.Google Scholar
Dibble, H. L., Abodolahzadeh, A., et al. 2017. How did hominins adapt to Ice Age Europe without fire? Current Anthropology 58:S278S287.Google Scholar
Dibble, H. L., Aldeias, V., et al. 2013. On the industrial attributions of the Aterian and Mousterian of the Maghreb. Journal of Human Evolution 64:194210.CrossRefGoogle ScholarPubMed
Dillehay, T. H. 1997. Monte Verde: A Late Pleistocene Settlement in Chile: The Archaeological Context, vol. 2. Washington, DC: Smithsonian Institution Press.Google Scholar
Dirks, P. H. G. M., Berger, L. R., et al. 2015. Geological and taphonomic context for the new hominin species Homo naledi from the Dinaledi Chamber, South Africa. eLife 4:e09561.Google Scholar
Dixon, E. J. 1999. Bones, Boats and Bison: Archeology and the First Colonization of Western North America. Albuquerque: University of New Mexico Press.Google Scholar
Djakovic, I., Key, A., and Soressi, M.. 2022. Optimal linear estimation models predict 1400–2900 years of overlap between Homo sapiens and Neandertals prior to their disappearance from France and northern Spain. Scientific Reports 12 (1):15000.Google Scholar
Druett, J. 2007. Island of the Lost: An Extraordinary Story of Survival at the Edge of the World. New York.: Algonquin Books.Google Scholar
Duke, H., and Pargeter, J.. 2015. Weaving simple solutions to complex problems: An experimental study of skill in bipolar cobble-splitting. Lithic Technology 40:349365.CrossRefGoogle Scholar
Efferson, C., Lalive, R., et al. 2008. The coevolution of cultural groups and ingroup favoritism. Science 321:18441849.Google Scholar
Ehrlich, P. R. 1968. The Population Bomb. New York: Sierra Club/Ballantine.Google Scholar
El Zaatari, S., Grine, F. E., et al. 2011. Ecogeographic variation in Neandertal dietary habits: Evidence from occlusal molar microwear texture analysis. Journal of Human Evolution 61:411424.Google Scholar
Emory, K. P. 1943. South Sea Lore. Honolulu: Bishop Museum (Special Publication 33).Google Scholar
Enzel, Y., and Bar-Yosef, O.. Editors. 2017. Quaternary of the Levant: Environments, Climate Change, and Humans. New York: Cambridge University Press.CrossRefGoogle Scholar
Eren, M. I., Lycett, S. J., et al. 2016. Test, model, and method validation: The role of experimental stone artifact replication in hypothesis-driven archaeology. Ethnoarchaeology 8:103136.Google Scholar
Erlandson, J. M. 2001. The archaeology of aquatic adaptations: Paradigms for a new millennium. Journal of Archaeological Research 9:287350.Google Scholar
Erlandson, J. M., Graham, M. H., et al. 2007. The Kelp Highway hypothesis: Marine ecology, the coastal migration theory, and the peopling of the Americas. The Journal of Island and Coastal Archaeology 2:161174.Google Scholar
Fagan, B. 2011. Cro-Magnon: How the Ice Age Gave Birth to the First Modern Humans. New York: Bloomsbury Press.Google Scholar
Faure, H., Walter, R. C., et al. 2002. The coastal oasis: Ice Age springs on emerged continental shelves. Global and Planetary Change 33:4756.Google Scholar
Feblot-Augustins, J. 1997. La circulation des matières premiers au Paléolithique: Synthése de données, perspectives comportementales. Liège: Etudes et Recherches Archéologiques de l’Université de Liège, No. 75.Google Scholar
Fedele, F., Giaccio, B, et al. 2004. The Campignian Ignimbrite eruption, Heinrich Event 4, and Palaeolithic Cange in Europe: A high-resolution investigation. Volcanism and the Earth’s Atmosphere (American Geophysical Union Geophysical Monograph) 139:301325.Google Scholar
Feidel, S. J. 2005. Man’s best friend – Mammoth’s worst enemy? A speculative essay on the role of dogs in Paleoindian colonization and megafaunal extinction. World Archaeology 37.Google Scholar
Fellows Yates, J. A., Velsko, I. M., et al. 2021. The evolution and changing ecology of the African hominid oral microbiome. Proceedings of the National Academy of Sciences 118:e2021655118.Google Scholar
Finlayson, C. 2003. Neanderthals and Modern Humans: The Dynamics and Biogeography of Their Interactions. Cambridge: Cambridge University Press.Google Scholar
Finlayson, C. 2019. The Smart Neanderthal: Cave Art, Bird Catching, and the Cognitive Revolution. New York: Oxford University Press.Google Scholar
Finney, B. 1979. Hokulea: The Way to Tahiti. New York: Dodd, Mead.Google Scholar
Firestone, R. B., West, A., et al. 2007. Evidence for an extraterrestrial impact 12,900 years ago that contributed to the megafaunal extinctions and the Younger Dryas cooling. Proceedings of the National Academy of Sciences 104:16016.Google Scholar
Flint, R. F. 1971. Glacial and Quaternary Geology. New York: John Wiley & Sons.Google Scholar
Flores, D. 2016. Coyote America: A Natural and Supernatural History. New York: Basic Books.Google Scholar
Florin, S. A., Fairbairn, A. S., et al. 2020. The first Australian plant foods at Madjedbebe, 65,000–53,000 years ago. Nature Communications 11:924.Google Scholar
Foley, R., and Lahr, M. M.. 1997. Mode 3 technologies and the evolution of modern humans. Cambridge Archaeological Journal 7:336.Google Scholar
Frison, G. C., and Bradley, B. A.. 1999. The Fenn Cache: Clovis Weapons & Tools. Santa Fe, NM: One Horse Land and Cattle Company.Google Scholar
Froehle, A. W., and Churchill, S. E.. 2009. Energetic competition between Neandertals and anatomically modern humans. PaleoAnthropology 2009:96116.Google Scholar
Fuller, D., and Hildebrand, E.. 2013. “Domesticating plants in Africa,” in The Oxford Handbook of African Archaeology. Edited by Mitchell, P. and Lane, P., pp. 131144. Oxford: Oxford University Press.Google Scholar
Gabriele, M., and Perry, D. M.. 2021. The Bright Ages: A New History of Medieval Europe. New York: Harper.Google Scholar
Gamble, C. 1987. “Man the Shoveler: Alternative Models for Middle Pleistocene Colonization and Occupation in Northern Latitudes,” in The Pleistocene Old World: Regional Perspectives. Edited by Soffer, O., pp. 8198. New York: Plenum.Google Scholar
Gamble, C. 2013. Settling the Earth: The Archaeology of Deep Human History. New York: Cambridge University Press.CrossRefGoogle Scholar
Gamble, C., Gowlett, J., et al. 2014. Thinking Big: How the Evolution of Social Life Shaped the Human Mind. New York: Thames and Hudson.Google Scholar
Garrod, D. A. E., and Bate, D. M. A.. Editors. 1937. The Stone Age of Mount Carmel, vol. 1: Excavations in the Wady el-Mughara. Oxford: Clarendon Press.Google Scholar
Geist, V. 1981. Neanderthal the hunter. Natural History 90:2636.Google Scholar
Gilligan, I. 2010. The prehistoric development of clothing: Archaeological implications of a thermal model. Journal of Archaeological Method and Theory 17:1580.Google Scholar
Gladikh, M. I., Kornietz, N., et al. 1984. Mammoth-bone dwellings on the Russian plain. Scientific American 251:164175.Google Scholar
Golding, W. 1954. The Lord of the Flies. London: Faber and Faber.Google Scholar
Gonzales, L. 2003. Deep Survival: Who Lives, Who Dies, and Why. New York: Norton.Google Scholar
Goodall, J. 1986. The Chimpanzees of Gombe: Patterns of Behavior. Cambridge, MA: Harvard University Press.Google Scholar
Goodwin, A. J. H. 1929. The Stone Ages in South Africa. Africa 2:174182.Google Scholar
Goren-Inbar, N. Editor. 1990. Quneitra: A Mousterian Site on the Golan Heights. Jerusalem: Hebrew University Institute of Archaeology (Qedem No. 31).Google Scholar
Gott, B. 2002. Fire-making in Tasmania: Absence of evidence is not evidence of absence. Current Anthropology 43:650655.Google Scholar
Gould, R. A. 1980. Living Archaeology. Cambridge: Cambridge University Press.Google Scholar
Graeber, D., and Wengrow, D.. 2021. The Dawn of Everything: A New History of Humanity. New York: Farrar, Straus & Giroux.Google Scholar
Greaves, R. D. 1997. “Hunting and Multifunctional Use of Bows and Arrows: Ethnoarchaeology of Technological Organization among the Pumé Hunters of Venezuela,” in Projectile Technology. Edited by Knecht, H., pp. 287320. New York: Plenum.Google Scholar
Green, R. E., Krause, J., et al. 2010. A draft sequence of the Neandertal genome. Science 328:710.CrossRefGoogle ScholarPubMed
Grinker, R. R. 1990. Images of denigration: Structuring inequality between foragers and farmers in the Ituri Forest, Zaire. American Ethnologist 17:111130.Google Scholar
Groube, L., Chappell, J., et al. 1986. A 40,000-year-old human occupation site at Huon Peninsula, Papua New Guinea. Nature 324:453455.Google Scholar
Groucutt, H. S., Scerri, E. M. L., et al. 2015. Stone tool assemblages and models for the dispersal of Homo sapiens out of Africa. Quaternary International 382:830.Google Scholar
Groucutt, H. S., White, T. S., et al. 2021. Multiple hominin dispersals into Southwest Asia over the past 400,000 years. Nature 597:376380.Google Scholar
Guthrie, R. D. 1990. The Frozen Fauna of the Mammoth Steppe: The Story of Blue Babe. Chicago, IL: University of Chicago Press.Google Scholar
Guthrie, R. D. 2001. Origin and causes of the mammoth steppe: A story of cloud cover, woolly mammal tooth pits, buckles, and inside-out Beringia. Quaternary Science Review 20:549574.CrossRefGoogle Scholar
Guthrie, R. D. 2005. The Nature of Paleolithic Art. Chicago, IL: University of Chicago Press.Google Scholar
Habgood, P. J., and Franklin, N. R.. 2008. The revolution that didn’t arrive: A review of Pleistocene Sahul. Journal of Human Evolution 55:187222.CrossRefGoogle Scholar
Hallinan, E., Barzilai, O., et al. 2022. No direct evidence for the presence of Nubian Levallois technology and its association with Neanderthals at Shukbah Cave. Science Reports:1204.Google Scholar
Hammond, G., and Hammond, N.. 1981. Child’s play: A distorting factor in archaeological distribution. American Antiquity 46:634636.Google Scholar
Hardy, B. L., Moncel, M. H., et al. 2020. Direct evidence of Neanderthal fibre technology and its cognitive and behavioral implications. Scientific Reports 10:4889.Google Scholar
Harmand, S., Lewis, J. E., et al. 2015. 3.3-million-year-old stone tools from Lomekwi 3, West Turkana, Kenya. Nature 521:310315.CrossRefGoogle ScholarPubMed
Harris, M. 2013. The Rise of Anthropological Theory, 4th ed. Toronto: University of Toronto Press.Google Scholar
Harvati, K., Stringer, C., et al. 2011. The Later Stone Age Calvaria from Iwo Eleru, Nigeria: Morphology and chronology. PLoS ONE 6:e24024.Google Scholar
Hayden, B. 1979. Palaeolithic Reflections: Lithic Technology and Ethnographic Excavations among Australian Aborigines. Canberra: Australian Institute of Aboriginal Studies.Google Scholar
Haynes, C. V., and Hemmings, E. T.. 1968. Mammoth-bone shaft wrench from Murray Springs, Arizona. Science 159:186187.Google Scholar
Heinrich, H. 1988. Origin and consequences of cyclic ice rafting in the Northeast Atlantic Ocean during the past 130,000 years. Quaternary Research 29:142152.Google Scholar
Henrich, J. 2004. Demography and cultural evolution: How adaptive cultural processes can produce maladaptive losses – The Tasmanian case. American Antiquity 69:197214.Google Scholar
Henry, A. G., Brooks, A. S., et al. 2014. Plant foods and the dietary ecology of Neanderthals and early modern humans. Journal of Human Evolution 69:4454.Google Scholar
Henshilwood, C., d’Errico, F., et al. 2004. Middle Stone Age shell beads from South Africa. Science 304:404.Google Scholar
Henshilwood, C. S., and Marean, C. W.. 2003. The origin of modern human behavior. Current Anthropology 44:627651.Google Scholar
Hershkovitz, I., May, H., et al. 2021. A Middle Pleistocene Homo from Nesher Ramla, Israel. Science 372:14241428.Google Scholar
Heyes, P. J., Anastasakis, K., et al. 2016. Selection and use of manganese dioxide by Neanderthals. Scientific Reports 6:22159.Google Scholar
Higham, C. 2014. Early Mainland Southeast Asia: From First Humans to Angkor. London: River Books Press.Google Scholar
Higham, T. 2021. The World before Us: The New Science behind Our Human Origins. New Haven, CT: Yale University Press.Google Scholar
Higham, T., Douka, K., et al. 2014. The timing and spatiotemporal patterning of Neanderthal disappearance. Nature 512:306.Google Scholar
Hiscock, P., O’Connor, S., et al. 2016. World’s earliest ground-edge axe production coincides with human colonisation of Australia. Australian Archaeology 82:211.Google Scholar
Hodder, I. 1982. Symbols in Action: Ethnoarchaeological Studies of Material Culture. Cambridge: Cambridge University Press.Google Scholar
Hoffecker, J. F. 2004. A Prehistory of the North: Human Settlement of the Higher Latitudes. Piscataway, NJ: Rutgers University Press.Google Scholar
Hoffecker, J. F. 2017. Modern Humans: Their African Origin and Global Dispersal. New York: Columbia University Press.Google Scholar
Holdaway, S., and Stern, N.. 2004. A Record in Stone: The Study of Australia’s Flaked Stone Artifacts. Melbourne: Museum Victoria/Aboriginal Studies Press.Google Scholar
Holldobler, B., and Wilson, E. O.. 1990. The Ants. Cambridge, MA: Belknap Press.Google Scholar
Hollenbach, K. D. 2007. “Gathering in the Late Paleoindian Period Archaeobotanical Remains from Dust Cave, Alabama,” in Foragers of the Terminal Pleistocene in North America. Edited by Walker, R. B and Driskell, B. N., pp. 132147. Lincoln: University of Nebraska Press.Google Scholar
Holliday, T. W. 2003. Species concepts, reticulation, and human evolution. Current Anthropology 44:653673.Google Scholar
Holliday, T. W. 2023. Cro-Magnon: The Story of the Last Ice Age People of Europe. New York: Columbia University Press.Google Scholar
Holliday, V. T., and Killick, D.. 2013. An early Paleoindian bead from the Mockingbird Gap Site, New Mexico. Current Anthropology 54:8595.Google Scholar
Holt, B. M., and Formicola, V.. 2008. Hunters of the Ice Age: The biology of Upper Paleolithic people. Yearbook of Physical Anthropology 15:7099.Google Scholar
Hovers, E., and Braun, D.. Editors. 2007. Interdisciplinary Approaches to the Oldowan. New York: Elsevier.Google Scholar
Hovers, E., Iliani, S., et al. 2003. An early case of color symbolism: Ochre use by modern humans in Qafzeh Cave. Current Anthropology 44:491522.CrossRefGoogle Scholar
Howell, F. C. 1952. Pleistocene glacial ecology and the evolution of “classic Neanderthal” man. Southwest Journal of Anthropology 8:377410.Google Scholar
Howell, F. C. 1959. Upper Pleistocene stratigraphy and early man in the Levant. Proceedings of the American Philosophical Society 103:165.Google Scholar
Howell, F. C., Butzer, K. W., et al. 1995. Observations on the Acheulean occupation site of Ambrona (Soria Province, Spain). Jahrbuch des Römisch-Germanischen Zentralmuseums Mainz 38:3382.Google Scholar
Hublin, J.-J. 2021. How old are the oldest Homo sapiens in Far East Asia? Proceedings of the National Academy of Sciences 118:e2101173118.CrossRefGoogle ScholarPubMed
Hume, D. 2018. Fire Making: The Forgotten Art of Conjuring Flame with Spark, Tinder, and Skil. New York: The Experiment.Google Scholar
Husain, A. 2018. The Sentient Machine: The Coming of Artificial Intelligence. New York: Scribner/Simon and Schuster.Google Scholar
Inskeep, R. 2001. “Some Notes on Fish and Fishing in Africa,” in A Very Remote Period Indeed: Papers on the Palaeolithic Presented to Derek Roe. Edited by Milliken, S. and Cook, J., pp. 6373. Oxford: Oxbow Books.Google Scholar
Ipek, H. 2009. Comparing and contrasting first and second language acquisition: Implications for language teachers. English Language Teaching 2:155162.Google Scholar
Isaac, G. L. 1981. Archaeological tests of alternative models of early hominid behaviour: Excavation and experiments. Philosophical Transactions of the Royal Society of London Series B, 292:177188.Google Scholar
Jasanoff, M. 2022. Ancestor worship: Where does the craze for geneaology come from? The New Yorker (May 2):6973.Google Scholar
Jelinek, A. J. Editor. 2013. Neandertal Lithic Industries at La Quina. Tucson: University of Arizona Press.Google Scholar
Johnson, C. R., and McBrearty, S.. 2010. 500,000 year old blades from the Kapthurin Formation, Kenya. Journal of Human Evolution 58:193200.CrossRefGoogle Scholar
Kaboth-Bahr, S., Gosling, W. D., et al. 2021. Paleo-ENSO influence on African environments and early modern humans. Proceedings of the National Academy of Sciences 118:e2018277118.Google Scholar
Keeley, L. H. 1996. War before Civilization. New York: Oxford University Press.Google Scholar
Kelly, R. L. 1998. “Foraging and Sedentism,” in Seasonality and Sedentism: Archaeological Perspectives from Old and New World Sites. Edited by Rocek, T. and Bar-Yosef, O., pp. 923. Cambridge, MA: Peabody Museum of Archaeology and Ethnology, Harvard University.Google Scholar
Kelly, R. L. 2013. The Lifeways of Hunter-Gatherers: The Foraging Spectrum, 2nd revised ed. New York: Cambridge University Press.Google Scholar
Kelton, P. 2007. Epidemics and Enslavement: Biological Catastrophe in the Native Southeast, 1492–1715. Lincoln: University of Nebraska Press.Google Scholar
Key, A., Fisch, M. R., et al. 2018. Early stage blunting causes rapid reductions in stone tool performance. Journal of Archaeological Science 91:111.Google Scholar
King, C. 2019. Gods of the Upper Air: How a Circle of Renegade Anthropologists Reinvented Race, Sex, and Gender in the Twentieth Century. New York: Anchor.Google Scholar
Kirch, P. V. 2018. On the Road of the Winds: An Archaeological History of the Pacific Islands before European Contact, 2nd ed. Berkeley: University of California Press.Google Scholar
Klein, R. G. 1992. The archaeology of modern human origins. Evolutionary Anthropology 1:514.Google Scholar
Kochanski, M. 1987. Bushcraft: Outdoor Skills and Wilderness Survival. Vancouver: Lone Pine Publishing.Google Scholar
Koester, R. J. 2008. Lost Person Behavior: A Search and Rescue Guide on Where to Look – For Land, Air and Water. Charlottesville, VA: dbS ProductionsGoogle Scholar
Kossina, G. 1911. Herkunft der Germanen. Leipzig: Kabitzich.Google Scholar
Kramer, K. 2005. Maya Children: Helpers at the Farm. Cambridge, MA: Harvard University Press.Google Scholar
Kramer, K. L. 2019. How there got to be so many of us: The evolutionary story of population growth and a life history of cooperation. Journal of Anthropological Research 75:472497.Google Scholar
Krauss, L. M. 2006. The Physics of Star Trek. New York: Basic Books.Google Scholar
Kroeber, A. L., and Kluckholn, C.. 1952. Culture; a Critical Review of Concepts and Definitions. Vol. 47, No. 1. Peabody Museum of American Archaeology and Ethnology. Cambridge, MA: Peabody Museum of American Archaeology and Ethnology, Harvard University.Google Scholar
Kuhn, S. L., and Clark, A. E.. 2015. Artifact densities and assemblage formation: Evidence from Tabun Cave. Journal of Anthropological Archaeology 38:816.CrossRefGoogle Scholar
Kuhn, S. L., and Stiner, M. C.. 2006. What’s a mother to do: A hypothesis about the division of labor among Neandertals and modern humans. Current Anthropology 47:953980.Google Scholar
Kuhn, S. L., and Stiner, M. C.. 2007. Paleolithic ornaments: Implications for cognition, demography and identity. Diogenes 54:4048.Google Scholar
Kuper, A. 1988. The Invention of Primitive Society: Transformations of an Illusion. London: Routledge.Google Scholar
Lahr, M. M., and Foley, R.. 1994. Multiple dispersals and modern human origins. Evolutionary Anthropology 3:4860.CrossRefGoogle Scholar
Laland, K. N. 2018. An evolved uniqueness: How we became a different kind of animal. Scientific American 319 (August):3239.Google Scholar
Landau, M. L. 1991. Narratives of Human Evolution. New Haven, CT: Yale University Press.Google Scholar
Langley, M. C., Amano, N., et al. 2020. Bows and arrows and complex symbolic displays 48,000 years ago in the South Asian tropics. Science Advances 6:eaba3831.Google Scholar
Langley, M. C., Clarkson, C., et al. 2019. Symbolic expression in Pleistocene Sahul, Sunda, and Wallacea. Quaternary Science Reviews 221:105883.Google Scholar
Larsen-Peterkin, G. 1993. “Lithic and Organic Hunting Technology in the French Upper Paleolithic,” in Hunting and Animal Exploitation in the Later Paleolithic and Mesolithic of Eurasia. Edited by Peterkin, G. L., Bricker, H., and Mellars, P. A., pp. 4968. Washington, DC: American Anthropological Association (Archaeological Paper No. 4).Google Scholar
Laville, H., Rigaud, J.-P., et al. 1980. Rock Shelters of the Perigord: Geological Stratigraphy and Archaeological Succession. New York: Academic Press.Google Scholar
Leakey, L. S. B. 1954. Adam’s Ancestors. London: Methuen.Google Scholar
Leakey, M. D. 1971. Olduvai Gorge: Excavations in Beds I and II, 1960–1963. Cambridge: Cambridge University Press.Google Scholar
Lee, R. B. 1979. The !Kung San: Men, Women and Work in a Foraging Society. Cambridge: Cambridge University Press.Google Scholar
Lepre, C. J., Roche, H., et al. 2011. An earlier origin for the Acheulian. Nature 477:8285.Google Scholar
Leslie, S., Winney, B., et al. 2015. The fine-scale genetic structure of the British population. Nature 519:309314.Google Scholar
Lesnik, Julie J. 2018. Edible Insects and Human Evolution. Gainesville, FL: University of Florida Press.Google Scholar
Levy, T. E. Editor. 1995. Archaeology of Society in the Holy Land. New York: Facts on File.Google Scholar
Lewis-Kraus, G. 2019. Game of bones: Is ancient DNA research revealing new truths – Or falling into old traps? New York Times Sunday Magazine (January 17):44.Google Scholar
Li, S., Schlebusch, C., et al. 2014. Genetic variation reveals large-scale population expansion and migration during the expansion of Bantu-speaking peoples. Proceedings of the Royal Society B: Biological Sciences 281:20141448.Google Scholar
Liebenberg, L. 1990. The Art of Tracking: The Origin of Science. Cape Town: David Phillip Publishers.Google Scholar
Lieberman, D. E. 2011. The Evolution of the Human Head. Cambridge, MA: Belknap/Harvard University Press.Google Scholar
Lieberman, D. E. 2013. The Story of the Human Body: Evolution, Health and Disease. New York: Pantheon Books.Google Scholar
Lieberman, D. E., and Shea, J. J.. 1994. Behavioral differences between archaic and modern humans in the Levantine Mousterian. American Anthropologist 96:300332.Google Scholar
Lieberman, P. 1973. On the Origins of Language. New York: Macmillan.Google Scholar
Lieberman, P. 2013. The Unpredictable Species: What Makes Humans Unique. Princeton, NJ: Princeton University Press.Google Scholar
Lieberman, P., and McCarthy, R. C.. 2014. “The Evolution of Speech and Language,” in Handbook of Paleoanthropology. Edited by Henke, W. and Tattersall, I., pp. 133. Berlin: Springer.Google Scholar
Lindly, J. M. 2005. The Mousterian of the Zagros: A Regional Perspective. Tempe: Arizona State University (Anthropological Research Papers, No. 56).Google Scholar
Lombard, M. 2020. The tip cross-sectional areas of poisoned bone arrowheads from southern Africa. Evolutionary Anthropology 33(5):307315.Google Scholar
Lombard, M. 2021. Variation in hunting weaponry for more than 300,000 years: A tip cross-sectional area study of Middle Stone Age points from southern Africa. Quaternary Science Reviews 264:107021.Google Scholar
Lombard, M., and Shea, J. J.. 2021. Did Pleistocene Africans use the spearthrower-and-dart? Evolutionary Anthropology 30:307315.Google Scholar
Lundin, C. 2003. 98.6 Degrees: The Art of Keeping Your Ass Alive. Salt Lake City, UT: Gibbs Smith.Google Scholar
Lundin, C. 2007. When All Hell Breaks Loose: Stuff You Need to Survive When Disaster Strikes. Salt Lake City, UT: Gibbs Smith.Google Scholar
Lycett, S. J., and Bae, C. J.. 2010. The Movius Line controversy: The state of the debate. World Archaeology 42:521544.Google Scholar
Lyell, C. 1830–1833. Principles of Geology: Being an Attempt to Explain the Former Changes of the Earth’s Surface, by Reference to Causes Now in Operation. London: John Murray.Google Scholar
Lyman, R. L. 1994. Vertebrate Taphonomy. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Lynas, M. 2009. Six Degrees: Our Future on a Hotter Planet. Washington, DC: National Geographic.Google Scholar
Maclaren, P. I. R. 1958. The Fishing Devices of Central and Southern Africa. Occasional Papers of the Rhodes-Livingstone Museum, No. 12. Livingstone: Rhodes-Livingstone Museum.Google Scholar
MacPhee, R. D. E. 2018. End of the Megafauna: The Fate of the World’s Hugest, Fiercest, and Strangest Animals. New York: W. W. Norton.Google Scholar
Maier, M. A., Barchfeld, P., et al. 2009. Context specificity of implicit preferences: The case of human preference for red. Emotion 9:734738.Google Scholar
Mann, C. C. 2011. 1493: Uncovering the New World Columbus Created. New York: Knopf.Google Scholar
Mann, C. C. 2018. The Wizard and the Prophet: Two Remarkable Scientists and Their Dueling Visions to Shape Tomorrow’s World. New York: Knopf.Google Scholar
Marean, C. W. 1991. Measuring the post-depositional destruction of bone in archaeological assemblages. Journal of Archaeological Science 18:677694.CrossRefGoogle Scholar
Marean, C. W. 2010. When the sea saved humanity. Scientific American 303:5561.Google Scholar
Marean, C. W. 2014. The origins and significance of coastal resource use in Africa and Western Eurasia. Journal of Human Evolution 77:1740.Google Scholar
Marean, C. W. 2015. An evolutionary anthropological perspective on modern human origins. Annual Review of Anthropology 44:533556.Google Scholar
Marean, C. W. 2016. The transition to foraging for dense and predictable resources and its impact on the evolution of modern humans. Philosophical Transactions of the Royal Society B: Biological Sciences 371.Google Scholar
Marlowe, F. W. 2010. The Hadza: Hunter-Gatherers of Tanzania. Berkeley: University of California Press.Google Scholar
Martin, P. S. 1973. The discovery of America: The first Americans may have swept the Western Hemisphere and decimated its fauna within 1000 years. Science 179:969974.Google Scholar
Marzke, M. W. 2013. Tool making, hand morphology and fossil hominins. Philosophical Transactions of the Royal Society B: Biological Sciences 368.Google Scholar
Marzke, M. W., and Wullstein, K. L.. 1996. Chimpanzee and human grips: A new classification with a focus on evolutionary morphology. International Journal of Primatology 17:117139.Google Scholar
Mayr, E. 2001. What Evolution Is. New York: Basic Books.Google Scholar
McBrearty, S., and Brooks, A. S.. 2000. The revolution that wasn’t: A new interpretation of the origin of modern human behavior. Journal of Human Evolution 39:453563.Google Scholar
McCown, T. D., and Keith, A.. 1939. The Stone Age of Mt. Carmel, vol. 2: The Fossil Human Remains from the Levalloiso-Mousterian. Oxford: Clarendon Press.Google Scholar
McGrew, W. C. 1992. Chimpanzee Material Culture: Implications for Human Evolution. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
McPherson, J., and McPherson, G.. 1993. Primitive Wilderness Living and Survival Skills. Randolph, KS: Prairie Wolf Publications.Google Scholar
McPherson, J., and McPherson, G.. 1996. Primitive Wilderness Living and Survival Skills: Applied and Advanced. Randolph, KS: Prairie Wolf Publications.Google Scholar
McPherron, S., Alemseged, Z., et al. 2010. Evidence for stone tool-assisted consumption of animal tissues before 3.39 million years ago at Dikika, Ethiopia. Nature 466:857860.Google Scholar
Mears, R. 2003. Essential Bushcraft: A Handbook of Survival Skills from around the World. London: Hodder and Stoughton.Google Scholar
Mellars, P. 1989a. Major issues in the emergence of modern humans. Current Anthropology 30:349385.CrossRefGoogle Scholar
Mellars, P. 1996a. The Neanderthal Legacy: An Archaeological Perspective from Western Europe. Princeton, NJ: Princeton University Press.Google Scholar
Mellars, P. 1996b. “Symbolism, Language, and the Early Human Mind,” in Modelling the Early Human Mind. Edited by Mellars, P. and Gibson, K., pp. 1532. Cambridge: MacDonald Institute for Archaeological Research.Google Scholar
Mellars, P. 2006. Going east: New genetic and archaeological perspectives on the modern human colonization of Eurasia. Science 313:796800.Google Scholar
Mellars, P. 2007. “Rethinking the Human Revolution: Eurasian and African Perspectives,” in Rethinking the Human Revolution. Edited by Mellars, P., Boyle, K., Bar-Yosef, O., and Stringer, C., pp. 114. Cambridge: McDonald Institute for Archaeological Research Monographs.Google Scholar
Mellars, P., and Stringer, C.. Editors. 1989. The Human Revolution: Behavioural and Biological Perspectives on the Origins of Modern Humans. Edinburgh: Edinburgh University Press.Google Scholar
Mellars, P. A. 1989b. “Technological Changes at the Middle-Upper Palaeolithic Transition: Economic, Social, and Cognitive Perspectives,” in The Human Revolution: Behavioural and Biological Perspectives on the Origins of Modern Humans. Edited by Mellars, P. and Stringer, C., pp. 338365. Edinburgh: Edinburgh University Press.Google Scholar
Mellars, P. A. Editor. 1990. The Emergence of Modern Humans: An Archaeological Perspective. Edinburgh: Edinburgh University Press.Google Scholar
Meltzer, D. J. 2021. First Peoples in a New World: Settling Ice Age America. Cambridge: Cambridge University Press.Google Scholar
Meltzer, D. J., Holliday, V. T., et al. 2014. Chronological evidence fails to support claim of an isochronous widespread layer of cosmic impact indicators dated to 12,800 years ago. Proceedings of the National Academy of Sciences 111:E2162.Google Scholar
Mercader, J., Barton, H., et al. 2007. 4,300-year-old chimpanzee sites and the origins of percussive stone technology. Proceedings of the National Academy of Sciences 104:30433048.CrossRefGoogle ScholarPubMed
Moffett, M. W. 2019. The Human Swarm: How Our Societies Arise, Thrive, and Fall. New York: Basic Books;.Google Scholar
Molnar, S., and Molnar, I. M.. 1985. The incidence of enamel hypoplasia among the Krapina Neandertals. American Anthropologist 87:536549.Google Scholar
Morcote-Ríos, G., Aceituno, F. J., et al. 2021. Colonisation and early peopling of the Colombian Amazon during the Late Pleistocene and the Early Holocene: New evidence from La Serranía La Lindosa. Quaternary International 578:519.Google Scholar
Movius, H. L., Jr. 1944. Early Man and Pleistocene Stratigraphy in Southeastern and Eastern Asia. Papers of the Peabody Museum Vol. 13, no. 3. Cambridge, MA.Google Scholar
Movius, H. L., Jr. 1950. A wooden spear of the Third Interglacial Age from Lower Saxony. Southwestern Journal of Anthropology 6:139142.Google Scholar
Muller, M. N., Wrangham, R. A., et al. 2017. Chimpanzees and Human Evolution. Cambridge, MA: Belknap Press (Harvard University Press).Google Scholar
Mulvaney, D. J. 1976. “The Chain of Connection: The Material Evidence,” in Tribes and Boundaries in Australia. Edited by Peterson, N., pp. 7294. Canberra: Australian Institute for Aboriginal Studies.Google Scholar
Nelson, N. 1916. “Flint-Working by Ishi,” in William Henry Holmes Anniversary Volume. Edited by Hodge, F. W., pp. 397402. Washington, DC: Smithsonian Institution.Google Scholar
Neves, W. A., Araujo, A. G. M., et al. 2012. Rock art at the Pleistocene/Holocene boundary in Eastern South America. PLoS ONE 7:e32228.Google Scholar
Newcomer, M. H. 1987. “Study and Replication of Bone Tools from Ksar Akil (Lebanon),” in Ksar Akil, Lebanon: A Technological and Typological Analysis of the Later Palaeolithic Levels of Ksar Akil, Vol. 2. Levels XIII–VI. Edited by Bergman, C. A., pp. 284307. British Archaeological Reports International Series, No. 329. Oxford: Oxbow Books.Google Scholar
Ní Leathlobhair, M., Perri, A. R., et al. 2018. The evolutionary history of dogs in the Americas. Science 361:8185.Google Scholar
Nicholson, S. L., Hosfield, R., et al. 2021. Beyond arrows on a map: The dynamics of Homo sapiens dispersal and occupation of Arabia during Marine Isotope Stage 5. Journal of Anthropological Archaeology 62:101269.Google Scholar
Niekus, M. J. L. T., Kozowyk, P. R. B., et al. 2019. Middle Paleolithic complex technology and a Neandertal tar-backed tool from the Dutch North Sea. Proceedings of the National Academy of Sciences of the United States of America 116:2208122087.Google Scholar
Novella, S. 2018. The Skeptic’s Guide to the Universe: How to Know What’s Really Real in a World Increasingly Fake. New York: Grand Central Publishing.Google Scholar
Novella, S., Novella, B., et al. 2022. The Skeptics’ Guide to the Future: What Yesterday’s Science and Science Fiction Tell Us about the World of Tomorrow. New York: Grand Central Publishing.Google Scholar
Nowell, A. 2010. Defining behavioral modernity in the context of Neandertal and anatomically modern human populations. Annual Review of Anthropology 39:437452.Google Scholar
Nowell, A. 2015. Learning to see and seeing to learn: Children, communities of practice and Pleistocene visual cultures. Cambridge Archaeological Journal 25:889899.Google Scholar
Nowell, A. 2021. Growing Up in the Ice Age: Fossil and Archaeological Evidence of the Lived Lives of Plio-Pleistocene Children. Oxford: Oxbow Books.CrossRefGoogle Scholar
O’Connell, J. F., Allen, J., et al. 2018. When did Homo sapiens first reach Southeast Asia and Sahul? Proceedings of the National Academy of Sciences 115:8482.Google Scholar
O’Connor, S., and Hiscock, P.. 2014. “The Peopling of Sahul and Near Oceania,” in The Oxford Handbook of Prehistoric Oceania. Edited by Cochrane, E. and Hunt, T., pp. 117. Oxford: Oxford University Press.Google Scholar
O’Connor, S., Ono, R., and Clarkson, C.. 2011. Pelagic fishing at 42,000 years before the present and the maritime skills of modern humans. Science 334 (6059):11171121.Google Scholar
Odell, G. H. 2004. Lithic Analysis. New York: Kluwer.CrossRefGoogle Scholar
Oliver, P. 1987. Dwellings: The House across the World. Dallas: University of Texas Press.Google Scholar
Oppenheimer, S. 2004. The Real Eve: Modern Man’s Journey Out of Africa. New York: Carroll & Graf.Google Scholar
Oswalt, W. H. 1973. Habitat and Technology: The Evolution of Hunting. New York: Holt, Rinehart and Winston.Google Scholar
Pargeter, J., and Eren, M.. 2017. Quantifying and comparing bipolar versus freehand flake morphologies, production currencies, and reduction energetics during lithic miniaturization. Lithic Technology 42(2–3):90108.Google Scholar
Pargeter, J., and Faith, J. T.. 2020. Lithic miniaturization as adaptive strategy: A case study from Boomplaas Cave, South Africa. Archaeological and Anthropological Sciences 12:225.Google Scholar
Pargeter, J. A., and Shea, J. J.. 2019. Going big vs. going small: Lithic miniaturization in hominin lithic technology. Evolutionary Anthropology 28:114.Google Scholar
Paulsen, G. 1986. The Hatchet. New York: Macmillan.Google Scholar
Péter, H., Zuberbühler, K., et al. 2022. Well-digging in a community of forest-living wild East African chimpanzees (Pan troglodytes schweinfurthii). Primates 63:355364.Google Scholar
Petraglia, M., Korisettar, R., et al. 2007. Middle Paleolithic assemblages from the Indian subcontinent before and after the Toba super-eruption. Science 317:114116.CrossRefGoogle ScholarPubMed
Phillipson, D. W. 1977. The Later Prehistory of Eastern and Southern Africa. London: Heinemann.Google Scholar
Phillipson, D. W. 2005. African Archaeology, 3rd ed. Cambridge: Cambridge University Press.Google Scholar
Pitts, M., and Roberts, M.. 1997. Fairweather Eden: Life in Britain Half a Million Years Ago as Revealed by the Excavations at Boxgrove. London: Century.Google Scholar
Pogue, D. 2021. How to Prepare for Climate Change: A Practical Guide to Surviving the Chaos. New York: Simon & Schuster.Google Scholar
Pope, G. G. 1989. Bamboo and human evolution. Natural History 1989:4857.Google Scholar
Pycraft, W. P., Elliot Smith, G., et al. 1928. Rhodesian Man and Associated Remains. London: British Museum (Natural History).Google Scholar
Radini, A., Buckley, S., et al. 2016. Neanderthals, trees and dental calculus: New evidence from El Sidrón. Antiquity 90:290301.Google Scholar
Rampino, M. R. 2017. Reexamining Lyell’s laws. American Scientist 1056:223231.Google Scholar
Rampino, M. R., Caldeira, K., et al. 2021. A pulse of the Earth: A 27.5-Myr underlying cycle in coordinated geological events over the last 260 Myr. Geoscience Frontiers 12:101245.CrossRefGoogle Scholar
Rampino, M. R., and Self, S.. 1992. Volcanic winter and accelerated glaciation following the Toba super-eruption. Nature 359:5052.Google Scholar
Reich, D. 2018. Who We Are and How We Got Here: Ancient DNA and the New Science of the Human Past. New York: Pantheon.Google Scholar
Reynolds, N., and Riede, F.. 2019. House of cards: Cultural taxonomy and the study of the European Upper Palaeolithic. Antiquity 93:13501358.Google Scholar
Richerson, P., Boyd, R., et al. 2001. Was agriculture impossible during the Pleistocene but mandatory during the Holocene? A climate change hypothesis. American Antiquity 66:387411.Google Scholar
Rigaud, A. 2001. Les bâtons percés: Décors énigmatiques et fonction possible. Gallia préhistoire 43:101151.Google Scholar
Rizal, Y., Westaway, K. E., et al. 2020. Last appearance of Homo erectus at Ngandong, Java, 117,000–108,000 years ago. Nature 577:381385.Google Scholar
Robinson, B. S., Ort, J. C., et al. 2009. Paleoindian aggregation and social context at Bull Brook. American Antiquity 74:423447.Google Scholar
Robinson, M. T. 2016. The Lost White Tribe: Explorers, Scientists, and the Theory That Changed a Continent. New York: Oxford University Press.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 108:52095214.Google Scholar
Rolland, N., and Dibble, H. L.. 1990. A new synthesis of Middle Paleolithic variability. American Antiquity 55:480499.Google Scholar
Rose, J. I. 2004. The question of Upper Pleistocene connections between East Africa and South Arabia. Current Anthropology 45:551555.Google Scholar
Rose, J. I., and Marks, A. E.. 2014. Out of Arabia and the Middle-Upper Paleolithic transition in the southern Levant. Quartär 61:4985.Google Scholar
Rose, J. I., Usik, V. I., et al. 2011. The Nubian complex of Dhofar, Oman: An African Middle Stone Age industry in southern Arabia. PLoS ONE 6:e28239.Google Scholar
Rots, V., and Plisson, H.. 2014. Projectiles and the abuse of the use-wear method in a search for impact. Journal of Archaeological Science 48:154165.Google Scholar
Ruditis, P. 2021. The Star Trek Book: Strange New Worlds Boldly Explained. New York: DK.Google Scholar
Sackett, J. R. 1991. “Straight Archaeology French Style: The Phylogenetic Paradigm in Historical Perspective,” in Perspectives on the Past: Theoretical Biases in Mediterranean Hunter-Gatherer Research. Edited by Clark, G., pp. 109139. Philadelphia: University of Pennsylvania Press.CrossRefGoogle Scholar
Saini, A. 2019. Superior: The Return of Race Science. Boston: Beacon Press.Google Scholar
Sandgathe, D. M., and Hayden, B.. 2003. Did Neanderthals eat inner bark? Antiquity 77:709718.CrossRefGoogle Scholar
Scerri, E. 2018. The origin of our species. New Scientist 238:3437.Google Scholar
Scerri, E. M. L. 2013. The Aterian and its place in the North African Middle Stone Age. Quaternary International 300:111130.Google Scholar
Schick, K. D., and Toth, N. P.. 1993. Making Silent Stones Speak: Human Evolution and the Dawn of Technology. New York: Simon and Schuster.Google Scholar
Schimelpfenig, T. 2016. National Outdoor Leadership School (NOLS) Wilderness Medicine, 6th ed. Mechanicsburg, PA: Stackpole Books for the National Outdoor Leadership School (NOLS).Google Scholar
Schmidt, P., Bellot-Gurlet, L., et al. 2018. The unique Solutrean laurel-leaf points of Volgu: Heat-treated or not? Antiquity 92:587602.Google Scholar
Schmidt, P., and Morala, A.. 2018. First insights into the technique used for heat treatment of chert at the Solutrean site of Laugerie-Haute, France. Archaeometry 60:885897.Google Scholar
Schutt, B. 2017. Cannibalism: A Perfectly Natural History. New York: Algonquin Books.Google Scholar
Serangeli, J., and Conard, N. J.. 2015. The behavioral and cultural stratigraphic contexts of the lithic assemblages from Schöningen. Journal of Human Evolution 89:287297.Google Scholar
Shackleton, N. J. 1987. Oxygen isotopes, ice volume, and sea level. Quaternary Science Reviews 6:183190.Google Scholar
Shackley, M. L. 1986. Still Living?: Yeti, Sasquatch and the Neanderthal Enigma. New York: W. W. Norton.Google Scholar
Shah, S. 2011. Fever: How Malaria Has Ruled Humankind for 500,000 Years. New York: Picador USA.Google Scholar
Shah, S. 2020. The Next Great Migration: The Beauty and Terror of Life on the Move. New York: Bloomsbury.Google Scholar
Shea, J. J. 1997. “Middle Paleolithic Spear Point Technology,” in Projectile Technology. Edited by Knecht, H., pp. 79106. New York: Plenum.Google Scholar
Shea, J. J. 2003. The Middle Paleolithic of the East Mediterranean Levant. Journal of World Prehistory 17:313394.Google Scholar
Shea, J. J. 2008a. The Middle Stone Age archaeology of the Lower Omo Valley Kibish Formation: Excavations, lithic assemblages, and inferred patterns of early Homo sapiens behavior. Journal of Human Evolution (Special Issue: Paleoanthropology of the Kibish Formation, Southern Ethiopia) 55:448485.Google Scholar
Shea, J. J. 2008b. Transitions or turnovers? Climatically-forced extinctions of Homo sapiens and Neanderthals in the East Mediterranean Levant. Quaternary Science Reviews 27:22532270.CrossRefGoogle Scholar
Shea, J. J. 2010. “Stone Age Visiting Cards Revisited: A Strategic Perspective on the Lithic Technology of Early Hominin Dispersal,” in Out of Africa 1: The First Hominin Colonization of Eurasia. Edited by Fleagle, J. G., Shea, J. J., Grine, F. E., Baden, A. L., and Leakey, R., pp. 4764. New York: Springer.Google Scholar
Shea, J. J. 2011a. Homo sapiens is as Homo sapiens was: Behavioral variability vs. “behavioral modernity” in Paleolithic archaeology. Current Anthropology 52:135.Google Scholar
Shea, J. J. 2011b. Refuting a myth of human origins. American Scientist 99:128135.Google Scholar
Shea, J. J. 2011c. “Sorting Out the Muddle in the Middle East: Glynn Isaac’s Method of Multiple Working Hypotheses Applied to Theories of Human Evolution in the Late Pleistocene Levant,” in Casting the Net Wide: Papers in Honor of Glynn Isaac and His Approach to Human Origins Research. Edited by Sept, J. M. and Pilbeam, D. R., pp. 213231. Cambridge, MA: American School of Prehistoric Research/Peabody Museum Press.Google Scholar
Shea, J. J. 2011d. Stone tool analysis and human evolution: Some advice from Uncle Screwtape. Evolutionary Anthropology 20:4853.Google Scholar
Shea, J. J. 2013a. Lithic Modes A–I: A new framework for describing global-scale variation in stone tool technology illustrated with evidence from the East Mediterranean Levant. Journal of Archaeological Method and Theory 20:151186.Google Scholar
Shea, J. J. 2013b. Stone Tools in the Paleolithic and Neolithic of the Near East: A Guide. New York: Cambridge University Press.CrossRefGoogle Scholar
Shea, J. J. 2014. Sink the Mousterian? Named stone tool industries (NASTIES) as obstacles to investigating hominin evolutionary relationships in the Later Middle Paleolithic Levant. Quaternary International 350:169179.Google Scholar
Shea, J. J. 2015. Making and using stone tools: Advice for learners and teachers and insights for archaeologists. Lithic Technology 40:231248.Google Scholar
Shea, J. J. 2017a. Occasional, obligatory, and habitual stone tool use in hominin evolution. Evolutionary Anthropology 26:200217.Google Scholar
Shea, J. J. 2017b. Stone Tools in Human Evolution: Behavioral Differences among Technological Primates. New York: Cambridge University Press.Google Scholar
Shea, J. J. 2020a. Prehistoric Stone Tools of Eastern Africa: A Guide. New York: Cambridge University Press.Google Scholar
Shea, J. J. 2020b. Survival Archaeology: A New Agenda for Prehistory’s Future. The Society for American Archaeology Record 20:1721, 66.Google Scholar
Shea, J. J., and Bar-Yosef, O.. 2005. Who were the Skhul/Qafzeh people? An archaeological perspective on Eurasia’s earliest modern humans. Journal of the Israel Prehistoric Society 35:449466.Google Scholar
Shea, J. J., and Sisk, M. L.. 2010. Complex projectile technology and Homo sapiens dispersal into Western Eurasia. PaleoAnthropology 2010:100122.Google Scholar
Shipman, P. 1981. The Life History of a Fossil: An Introduction to Taphonomy and Paleoecology. Cambridge, MA: Harvard University Press.Google Scholar
Shipman, P. 2015. The Invaders: How Humans and Their Dogs Drove Neanderthals to Extinction. Cambridge, MA: Belknap/Harvard University Press.Google Scholar
Shipman, P. 2021. Our Oldest Companions: The Story of the First Dogs. Cambridge, MA: Belknap Press of Harvard University Press.Google Scholar
Shostak, M. 1990. Nisa: The Life and Words of a !Kung Woman, 4th ed. Cambridge, MA: Harvard University Press.Google Scholar
Silverman, D. J. 2016. Thundersticks: Firearms and the Violent Transformation of Native America. Cambridge, MA: Belknap Press.Google Scholar
Simmons, A. 2012. Mediterranean island voyages. Science 338:895.Google Scholar
Singer, R., and Wymer, J.. 1982. The Middle Stone Age at Klasies River Mouth in South Africa. Chicago, IL: University of Chicago Press.Google Scholar
Slimak, L., Zanolli, C., et al. 2022. Modern human incursion into Neanderthal territories 54,000 years ago at Mandrin, France. Science Advances 8:eabj9496.Google Scholar
Smith, G. M., Ruebens, K., et al. 2019. Subsistence strategies throughout the African Middle Pleistocene: Faunal evidence for behavioral change and continuity across the Earlier to Middle Stone Age transition. Journal of Human Evolution 127:120.Google Scholar
Smith, P. E. L. 1964. The Solutrean culture. Scientific American 211:8694.Google Scholar
Smith, T. M., Austin, C., et al. 2018. Wintertime stress, nursing, and lead exposure in Neanderthal children. Science Advances 4:eaau9483.Google Scholar
Soffer, O. 1985. The Upper Paleolithic of the Central Russian Plain. Orlando, FL: Academic Press.CrossRefGoogle Scholar
Sorensen, A., Roebroeks, W., et al. 2014. Fire production in the deep past? The expedient strike-a-light model. Journal of Archaeological Science 42:476486.Google Scholar
Sorensen, A. C., Claud, E., et al. 2018. Neandertal fire-making technology inferred from microwear analysis. Scientific Reports 8:10065.Google Scholar
Sorensen, B. 2009. Energy use by Eem Neanderthals. Journal of Archaeological Science 36:22012205.Google Scholar
Sorenson, M. V., and Leonard, W. R.. 2001. Neandertal energetics and foraging efficiency. Journal of Human Evolution 40:483495.Google Scholar
Speth, J. D. 2010. The Paleoanthropology and Archaeology of Big-Game Hunting: Protein, Fat or Politics? New York: Springer.Google Scholar
Speth, J. D., and Clark, J. L.. 2006. Hunting and overhunting in the Levantine Late Middle Palaeolithic. Before Farming 3:Article 1.Google Scholar
Stanford, D. J., and Bradley, B. A.. 2012. Across Atlantic Ice: The Origins of America’s Clovis Culture. Berkeley: University of California Press.Google Scholar
Stern, N. 2009. “The Archaeological Signature of Behavioral Modernity: A Perspective from the Southern Periphery of the Modern Human Range,” in Transitions in Prehistory: Essays in Honor of Ofer Bar-Yosef. Edited by Shea, J. J. and Lieberman, D. E., pp. 255287. Oxford: Oxbow Books.Google Scholar
Stiner, M. C. 1993. Modern human origins: Faunal perspectives. Annual Review of Anthropology 22:5582.Google Scholar
Stiner, M. C., Munro, N. D., et al. 1999. Paleolithic population growth pulses evidenced by small animal exploitation. Science 283:190194.CrossRefGoogle ScholarPubMed
Strasser, T., Panagopoulou, E., et al. 2010. Stone Age seafaring in the Mediterranean: Evidence from the Plakias region for Lower Palaeolithic and Mesolithic habitation of Crete. Hesperia 79:145190.Google Scholar
Straus, L. G. 1995. The Upper Paleolithic of Europe: An overview. Evolutionary Anthropology 4:416.Google Scholar
Straus, L. G., Meltzer, D. J., et al. 2005. Ice Age Atlantis? Exploring the Solutrean-Clovis “connection.” World Archaeology 37:507532.CrossRefGoogle Scholar
Stringer, C. 2012. Lone Survivors: How We Came to Be the Only Humans on Earth. New York: Times Books/Henry Holt.Google Scholar
Sullivan, A. P. 1989. The technology of ceramic reuse: Formation processes and archaeological evidence. World Archaeology 21:101114.Google Scholar
Sykes, B. 2001. The Seven Daughters of Eve: The Science That Reveals Our Ancestry. New York: W. W. Norton.Google Scholar
Sykes, R. W. 2020. Kindred: Neanderthal Life, Love, Death and Art. New York: Bloomsbury Sigma.Google Scholar
Taborin, Y. 2003. “La mer et les premiers hommes modernes,” in Echanges et diffusion dans la préhistoire méditerranéenne. Edited by Vandermeersch, B., pp. 113122. Paris: Editions du Comité des travaux historiques et scientifiques.Google Scholar
Taylor, D. M., and Doria, J. R.. 1981. Self-serving and group-serving bias in attribution. Journal of Social Psychology 113:201211.Google Scholar
Taylor, W. W. 1948. A Study of Archaeology. Washington, DC: American Anthropological Association (Memoir 69).Google Scholar
Tchernov, E. 1988. “The Biogeographical History of the Southern Levant,” in The Zoogeography of Israel. Edited by Yom-Tov, Y. and Tchernov, E., pp. 401409. Dordrecht: Junk.Google Scholar
Templeton, A. 1999. Races: A genetic and evolutionary perspective. American Anthropologist 100:632650.Google Scholar
Tennie, C., Call, J., et al. 2009. Ratcheting up the ratchet: On the evolution of cumulative culture. Philosophical Transactions of the Royal Society B: Biological Sciences 364:24052415.Google Scholar
Texier, P.-J., Porraz, G., et al. 2010. A Howiesons Poort tradition of engraving ostrich eggshell containers dated to 60,000 years ago at Diepkloof Rock Shelter, South Africa. Proceedings of the National Academy of Sciences 107:61806185.Google Scholar
Theime, H. 1997. Lower Paleolithic hunting spears from Germany. Nature 385:807810.Google Scholar
Thomas, N. 2021. Voyagers: The Settlement of the Pacific. New York: Basic Books.Google Scholar
Thomson, D. F. 1939. The seasonal factor in human culture. Proceedings of the Prehistoric Society 10:209221.Google Scholar
Thurston, T. L., and Fisher, C. T.. 2007. “Seeking a Richer Harvest: An Introduction to the Archaeology of Subsistence Intensification, Innovation, and Change,” in Seeking a Richer Harvest: The Archaeology of Subsistence Intensification, Innovation, and Change. Edited by Thurston, T. L. and Fisher, C. T., pp. 123. New York: Springer.Google Scholar
Tillet, T. 1996. “Behaviour Patterns, Strategies, and Seasonality in the Mousterian site of Prélétang, (Vercours): The Mousterian of the Alps,” in Middle Palaeolithic and Middle Stone Age Settlement Systems. Edited by Conard, N. J. and Wendorf, F., pp. 319326. Forli: ABACO Edizione.Google Scholar
Tillier, A.-M. 2002. “Investigating the Biological Evidence for Stress and Disease in Levantine Early Anatomically Modern Humans.” Human Paleoecology in the Levantine Corridor, Institute for Advanced Studies, Hebrew University, Jerusalem, Israel, 2002, pp. 135–148.Google Scholar
Toups, M. A., Kitchen, A., et al. 2011. Origin of clothing lice indicates early clothing use by anatomically modern humans in Africa. Molecular Biology and Evolution 28:2932.CrossRefGoogle ScholarPubMed
Towell, C. 2020. The Survival Handbook, new ed. New York: DK Publishing.Google Scholar
Trauth, M. H., Maslin, M. A., et al. 2010. Human evolution in a variable environment: The amplifier lakes of Eastern Africa. Quaternary Science Reviews 29:29812988.CrossRefGoogle Scholar
Trigger, B. G. 2006. A History of Archaeological Thought, 2nd ed. Cambridge: Cambridge University Press.Google Scholar
Trinkaus, E. 1983. The Shanidar Neanderthals. New York: Academic Press.Google Scholar
Trinkaus, E. 1984. “Western Asia,” in The Origins of Modern Humans. Edited by Smith, F. H. and Spencer, F., pp. 251293. New York: Alan R. Liss.Google Scholar
Trinkaus, E. 2006. Modern human versus Neandertal evolutionary distinctiveness. Current Anthropology 47:597620.Google Scholar
Trinkaus, E., and Shipman, P.. 1993. The Neandertals: Changing the Image of Mankind. New York: Knopf.Google Scholar
Trinkaus, E., and Svoboda, J.. 2005. Early Modern Human Evolution in Central Europe: The People of Dolní Vestonice and Pavlov. New York: Oxford University Press.Google Scholar
Tryon, C. A., and Faith, J. T.. 2013. Variability in the Middle Stone Age of Eastern Africa. Current Anthropology 54 (S8):S234–S254.CrossRefGoogle Scholar
Tryon, C. A., Lewis, J. E., et al. 2018. Middle and Later Stone Age chronology of Kisese II rockshelter (UNESCO World Heritage Kondoa Rock-Art Sites), Tanzania. PLoS ONE 13:e0192029.Google Scholar
Twiss, K. C. 2019. The Archaeology of Food: Identity, Politics, and Ideology in the Prehistoric and Historic Past. New York: Cambridge University Press.Google Scholar
Vandiver, P. B., Soffer, O., et al. 1989. The origins of ceramic technology at Dolni Vƒõstonice, Czechoslovakia. Science 246:10021008.CrossRefGoogle ScholarPubMed
Vanhaeren, M., and d’Errico, F.. 2006. Aurignacian ethno-linguistic geography of Europe revealed by personal ornaments. Journal of Archaeological Science 33:11051128.Google Scholar
Vanhaeren, M., d’Errico, F., et al. 2006. Middle Paleolithic shell beads in Israel and Algeria. Science 312:17851788.Google Scholar
Wadley, L., Williamson, B., et al. 2004. Ochre in hafting in Middle Stone Age southern Africa: A practical role. Antiquity 78:661675.Google Scholar
Walker, R. B. 2007. “Hunting in the Late Paleoindian Period Faunal Remains from Dust Cave, Alabama,” in Foragers of the Terminal Pleistocene in North America. Edited by Walker, R. B. and Driskell, B. N., pp. 99115. Lincoln: University of Nebraska Press.CrossRefGoogle Scholar
Waters, M. R., Stafford, T. W, et al. 2020. The age of Clovis – 13,050 to 12,750 cal yr B.P. Science Advances 6:eaaz0455.Google Scholar
Wells, S. 2009. Deep Ancestry: Inside the Genographic Project. Washington, DC: National Geographic Society.Google Scholar
Werner, D., Thurman, C., et al. 2010. Where There Is No Doctor: A Village Health Care Handbook. Berkeley, CA: Hesperian.Google Scholar
White, J. P., and O’Connell, J. F.. 1982. A Prehistory of Australia, New Guinea and Sahul. New York: Academic Press.Google Scholar
White, R. 1986. Dark Caves, Bright Visions. New York: American Museum of Natural History.Google Scholar
Whiten, A., Goodall, J., et al. 1999. Cultures in chimpanzees. Nature 399:682685.Google Scholar
Whittaker, J. C. 1994. Flintknapping: Making and Understanding Stone Tools. Austin: University of Texas Press.Google Scholar
Whittaker, J. C. 2004. American Flintknappers: Stone Age Art in the Age of Computers. Austin: University of Texas Press.Google Scholar
Wilkins, J., and Chazan, M.. 2012. Blade production ∼500 thousand years ago at Kathu Pan 1, South Africa: Support for a multiple origins hypothesis for early Middle Pleistocene blade technologies. Journal of Archaeological Science 39:18831900.Google Scholar
Wilkins, J., Schoville, B. J., et al. 2012. Evidence for early hafted hunting technology. Science 338:942946.CrossRefGoogle ScholarPubMed
Will, M., Mackay, A., et al. 2015. Implications of Nubian-like core reduction systems in southern Africa for the identification of early modern human dispersals. PLoS ONE 10:e0131824.CrossRefGoogle ScholarPubMed
Willoughby, P. R. 2007. The Evolution of Modern Humans in Africa: A Comprehensive Guide. New York: Altamira.Google Scholar
Wilmsen, E. N. 1989. Land Filled with Flies: A Political Economy of the Kalahari. Chicago, IL: University of Chicago Press.Google Scholar
Wobst, H. M. 1974. Boundary conditions or paleolithic social systems: A simulation approach. American Antiquity 39:303309.Google Scholar
Wobst, M. H. 1977. “Stylistic Behavior and Information Exchange,” in For the Director, Research Essays in Honor of James B. Griffin. Edited by Cleland, C., pp. 317342. Ann Arbor: Museum of Anthropology, University of Michigan (Anthropology Paper No. 61).Google Scholar
Wolf, E. R. 1982. Europe and the People without History. Berkeley: University of California Press.Google Scholar
Wolpoff, M., and Thorne, A.. 2003. The multiregional evolution of humans. Scientific American 13:4653.Google Scholar
Wolpoff, M. H., and Caspari, R.. 1997. Race and Human Evolution: A Fatal Attraction. New York: Simon and Schuster.Google Scholar
Wrangham, R. W. 1999. Evolution of coalitionary killing. Yearbook of Physical Anthropology 42:130.Google Scholar
Wrangham, R. W. 2009. Catching Fire: How Cooking Made Us Human. New York: Basic Books.Google Scholar
Wrangham, R. W. 2018. Two types of aggression in human evolution. Proceedings of the National Academy of Sciences 115:245.Google Scholar
Wrangham, R. W. 2019. The Goodness Paradox: The Strange Relationship between Virtue and Violence in Human Evolution. New York: Pantheon.Google Scholar
Wynn, T., and Coolidge, F. L.. 2011. How to Think like a Neanderthal. Oxford: Oxford University Press.Google Scholar
Xia, L., Robock, A., et al. 2022. Global food insecurity and famine from reduced crop, marine fishery and livestock production due to climate disruption from nuclear war soot injection. Nature Food 3:586596.Google Scholar
Yaroshevich, A., Kaufman, D., et al. 2021. Weapons in transition: Reappraisal of the origin of complex projectiles in the Levant based on the Boker Tachtit stratigraphic sequence. Journal of Archaeological Science 131:105381.Google Scholar
Yellen, J. 1977. Archaeological Approaches to the Present: Models for Reconstructing the Past. New York: Academic Press.Google Scholar
Yellen, J., Brooks, A., et al. 2005. The archaeology of the Aduma Middle Stone Age sites in the Awash Valley, Ethiopia. PaleoAnthropology 10:25100.Google Scholar
Yellen, J. E., Brooks, A. S., et al. 1995. A Middle Stone Age Worked Bone Industry from Katanda, Upper Semliki Valley, Zaire. Science 268:553556.Google Scholar
Zahavi, A., and Zahavi, A.. 1997. The Handicap Principle: A Missing Piece of Darwin’s Puzzle. Oxford: Oxford University Press.CrossRefGoogle Scholar
Zaidner, Y., Centi, L., et al. 2021. Middle Pleistocene Homo behavior and culture at 140,000 to 120,000 years ago and interactions with Homo sapiens. Science 372:14291433.Google Scholar
Zilhao, J. 2011. “Aliens from Outer Time? Why the ‘Human Revolution’ Is Wrong, and Where Do We Go from Here?,” in Continuity and Discontinuity in the Peopling of Europe: One Hundred Fifty Years of Neanderthal Study. Edited by Condemi, S. and Weniger, G.-C., pp. 331366. New York: Springer.Google Scholar
Zilhão, J., Angelucci, D. E., et al. 2020. Last Interglacial Iberian Neandertals as fisher-hunter-gatherers. Science 367:eaaz7943.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.

  • Bibliography
  • John J. Shea, State University of New York, Stony Brook
  • Book: The Unstoppable Human Species
  • Online publication: 16 March 2023
  • Chapter DOI: https://doi.org/10.1017/9781108554060.020
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.

  • Bibliography
  • John J. Shea, State University of New York, Stony Brook
  • Book: The Unstoppable Human Species
  • Online publication: 16 March 2023
  • Chapter DOI: https://doi.org/10.1017/9781108554060.020
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.

  • Bibliography
  • John J. Shea, State University of New York, Stony Brook
  • Book: The Unstoppable Human Species
  • Online publication: 16 March 2023
  • Chapter DOI: https://doi.org/10.1017/9781108554060.020
Available formats
×