Skip to main content Accessibility help
×
Hostname: page-component-7479d7b7d-q6k6v Total loading time: 0 Render date: 2024-07-09T06:25:42.613Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  27 October 2021

Jennifer C. French
Affiliation:
University of Liverpool
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
Palaeolithic Europe
A Demographic and Social Prehistory
, pp. 275 - 326
Publisher: Cambridge University Press
Print publication year: 2021

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

Acerbi, A., Kendal, J. & Tehrani, J. J.. 2017. Cultural complexity and demography: the case of folktales. Evolution and Human Behavior 38 (4): 474–80.CrossRefGoogle Scholar
Agustí, J., Blain, H-A, Cuenca-Bescós, G. & Bailon, S.. 2009. Climate forcing of first hominid dispersal in Western Europe. Journal of Human Evolution 57: 815–21.Google Scholar
Aiello, L. C. & Key, C.. 2002. Energetic consequences of being a Homo erectus female. American Journal of Human Biology 14: 551–65.Google Scholar
Aiello, L. C. & Wheeler, P.. 2003. Neanderthal thermoregulation and the glacial climate. In van Andel, T. H & Davies, W. (eds.), Neanderthals and Modern Humans in the European Landscape during the Last Glaciation: Archaeological Results of the Stage 3 Project. Cambridge: McDonald Institute Monographs, pp. 147–66.Google Scholar
Alcaraz-Castaño, M. 2015. Central Iberia around the Last Glacial Maximum: hopes and prospects. Journal of Anthropological Research 71: 565–78.Google Scholar
Allain, J. & Fritsch, R.. 1967. Le Badegoulien de l’Abri Fritsch aux Roches de Pouligny-St-Pierre (Indre). Bulletin de la Société Préhistorique Française 64: 8394.Google Scholar
Allsworth-Jones, P. 1986. The Szeletian and the Transition from the Middle to Upper Palaeolithic in Central Europe. Oxford: Clarendon Press.Google Scholar
Álvarez, C., Parés, J. M., Granger, D., Duval, M., Sala, R. & Toro, I.. 2015. New magnetostratigraphic and numerical age of the Fuente Nueva-3 site (Guadix-Baza basin, Spain). Quaternary International 389: 224–34.Google Scholar
Alves, I., Arenas, M., Currat, M., A., et al. 2016. Long-distance dispersal shaped patterns of human genetic diversity in Eurasia. Molecular Biology and Evolution 33 (4): 946–58.Google Scholar
Anderson, L., Reynolds, N. & Teyssandier, N.. 2019. No reliable evidence for a very early Aurignacian in Southern Iberia. Nature Ecology & Evolution 3 (5): 713.Google Scholar
Anderson, R. M. & May, R. M.. 1991. Infectious Diseases of Humans: Dynamics and Control. Oxford: Oxford University Press.Google Scholar
Andrews, J. T. 1998. Abrupt changes (Heinrich events) in late Quaternary North Atlantic marine environments: a history and review of data and concepts. Journal of Quaternary Science 13 (1): 316.Google Scholar
Anghelinu, M., Niţă, L. & Murătoreanu, G.. 2018. Le Gravettien et l’Épigravettien de l’Est de la Roumanie: une réévaluation. L’Anthropologie 122 (2): 183219.Google Scholar
Antón, S. C., Potts, R. & Aiello, L. C.. 2014. Evolution of early Homo: an integrated biological perspective. Science 345 (6192): 1236828.Google Scholar
Aoki, K. 2018. On the absence of a correlation between population size and ‘toolkit size’ in ethnographic hunter-gatherers. Philosophical Transactions of the Royal Society B, 373: 20170061.Google Scholar
Apicella, C. L., Marlowe, F. W., Fowler, J. H. & Christakis, N. A.. 2012. Social networks and cooperation in hunter-gatherers. Nature 481: 497502.Google Scholar
Aranburu, A., Arsuaga, J. L. & Sala, N.. 2017. The stratigraphy of the Sima de los Huesos (Atapuerca, Spain) and implications for the origin of the fossil hominin accumulation. Quaternary International 433: 521.CrossRefGoogle Scholar
Arazello, M., Peretto, C., Moncel, M-H. 2015. The Pirro Nord site (Apricena, Fg, Southern Italy) in the context of the first European peopling: convergences and divergences. Quaternary International 389: 255–63.Google Scholar
Armelagos, G. J., Brown, P. J. & Turner, B.. 2005. Evolutionary, historical and political economic perspectives on health and disease. Social Science & Medicine 61: 755–65.CrossRefGoogle ScholarPubMed
Arrighi, S., Bortolini, E., Tassoni, L., et al. 2020. Backdating systematic shell ornament making in Europe to 45,000 years ago. Archaeological and Anthropological Sciences 12: 59.CrossRefGoogle Scholar
Arsuaga, J. L., Martínez, I., Arnold, L. J., et al. 2014. Neandertal roots: cranial and chronological evidence from Sima de los Huesos. Science 344 (6190): 1358–63.Google Scholar
Ashton., N. 2015. Ecological niches, technological developments and physical adaptations of early humans in Europe: the handaxe-Heidelbergensis hypothesis. In Coward, F, Hosfield, R, Pope, M & Wenban-Smith, F (eds.), Settlement, Society and Cognition in Human Evolution: Landscapes in Mind. Cambridge: Cambridge University Press, pp. 138–53.Google Scholar
Ashton, N. 2017. Landscapes of habit and persistent places during MIS 11 in Europe. In Pope, M., McNabb, J. & Gamble, C. (eds.), Crossing the Human Threshold: Dynamic Transformation and Persistent Places during the Middle Pleistocene. London: Routledge, pp. 142–64.Google Scholar
Ashton, N. & Hosfield, R.. 2009. Mapping the human record in the British early Palaeolithic: evidence from the Solent River system. Journal of Quaternary Science 25 (5): 737–53.Google Scholar
Ashton, N. & Lewis, S.. 2002. Deserted Britain: declining populations in the British Late Middle Pleistocene. Antiquity 76: 791798.CrossRefGoogle Scholar
Ashton, N. & Lewis, S. G.. 2012. The environmental contexts of early human occupation of northwest Europe: the British Lower Palaeolithic record. Quaternary International 271: 5064.Google Scholar
Ashton, N., Lewis, S. G., De Groote, I., et al. 2014. Hominin footprints from Early Pleistocene deposits at Happisburgh, UK. Plos One 9 (2): e88329.Google Scholar
Ashton, N., Lewis, S. G., Parfitt, S. A., Davis, R. J. & Stringer, C.. 2016. Handaxe and non-handaxe assemblages during Marine Isotope Stage 11 in northern Europe: recent investigations at Barnham, Suffolk, UK. Journal of Quaternary Science 31 (8): 837–43.CrossRefGoogle Scholar
Ashton, N., Harris, C. R. E. & Lewis, S. G.. 2018. Frontiers and route-ways from Europe: the Early Middle Palaeolithic of Britain. Journal of Quaternary Science 33 (2): 194211.Google Scholar
Attenbrow, V. 2006. What’s Changing: Population Size or Land-Use Patterns? The Archaeology of Upper Mangrove Creek, Sydney Basin (Terra Australis 21). Canberra: Pandanus Books.Google Scholar
Audouze, F. 2010. Domesticity and spatial organization at Verberie. In Zubrow, E., Audouze, F. & Enloe, J. (eds.), The Magdalenian Household: Unraveling Domesticity. Albany: State University of New York Press, pp. 145–75.Google Scholar
Aura Tortosa, J. E. 2007. Badegouliens et Magdaléniens du versant méditerranéen espagnol. Bulletin de la Société Préhistorique Française 104 (4): 809–24.Google Scholar
Austin, C., Smith, T. M., Bradman, A., et al. 2013. Barium distributions in teeth reveal early-life dietary transitions in primates. Nature 498: 216–20.Google Scholar
Baales, M. 2014. Untermassfeld- or the struggle for finding the earliest traces of human occupation in Central Europe: a comment on ‘Hominin dispersals from the Jaramillo subchron in central and south-western Europe: Untermassfeld (Germany) and Vallparadís (Spain)’ by Garcia, J. et al., Quaternary International 316 (2013), pp 7393. Quaternary International 337: 254–56Google Scholar
Baales, M., Jöris, O., Justus, A. & Roebroeks, W.. 2000. Natur oder Kultur? Zur Frage ältestpaläolithischer Artefaktensembles aus Hauptterrassenschottern in Deutschland. Germania: Anzeiger der Römisch-Germanischen Komission des Deutschen Archäologischen Institute 78 (1): 120.Google Scholar
Bachechi, L., Fabbri, P-F & Mallegni, F.. 1997. An arrow-caused lesion in a Late Upper Palaeolithic human pelvis. Current Anthropology 38 (1): 135–40.Google Scholar
Bae, C. J., Douka, K. & Petraglia, M. D.. 2017. On the origin of modern humans: Asian perspectives. Science 358: eaai9067.CrossRefGoogle ScholarPubMed
Bahn, P. G. 1982. Inter‐site and inter‐regional links during the Upper Palaeolithic: the Pyrenean evidence. Oxford Journal of Archaeology 1 (3): 247–68.Google Scholar
Bailey, R. C. & Aunger, R. V.. 1995. Sexuality, infertility and sexually transmitted disease among farmers and foragers in Central Africa. In Abramson, P. R. & Pinkerton, S. D. (eds.), Sexual Nature, Sexual Culture: Chicago: The University of Chicago Press, pp. 195222.Google Scholar
Bailey, S. E. & Hublin, J. J.. 2005. Who made the Early Aurignacian? A reconsideration of the Brassempouy dental remains. Bulletins et Mémoires de la Société d’Anthropologie de Paris 17 (1–2): 115–21.Google Scholar
Bailey, S. E., Weaver, T. D. & Hublin, J-J. 2009. Who made the Aurignacian and other early Upper Palaeolithic industries? Journal of Human Evolution 57: 1126.CrossRefGoogle Scholar
Balme, J. & Bowdler, S.. 2006. Spear and digging stick: the origin of gender and its implications for the colonization of new continents. Journal of Social Archaeology 6 (3): 379401.Google Scholar
Banks, W. E., d’Errico, F., Peterson, A. T, Kageyama, M, Sima, A & Sánchez-Goñi, M. F. 2008a. Neanderthal extinction by competitive exclusion. PLoS One 3 (12): e3972.Google Scholar
Banks, W. E., d’Errico, F, Peterson, A. T, Vanhaeren, M et al. 2008b. Human ecological niches and ranges during the LGM in Europe derived from an application of eco-cultural niche modelling. Journal of Archaeological Science 35: 481–91.CrossRefGoogle Scholar
Banks, W. E., d’Errico, F & Zilhão, J. 2013. Human–climate interaction during the Early Upper Paleolithic: testing the hypothesis of an adaptive shift between the Proto-Aurignacian and the Early Aurignacian. Journal of Human Evolution 64 (1): 3955.Google Scholar
Bard, E., Heaton, T. J., Talamo, S., Kromer, B., Reimer, R. W. & Reimer, P. J.. 2020. Extended dilation of the radiocarbon time scale between 40,000 and 48,000 y BP and the overlap between Neanderthals and Homo sapiens. Proceedings of the National Academy of Sciences USA 117 (35): 21005–7.CrossRefGoogle Scholar
Barrett, C. K., Guatelli‐Steinberg, D. & Sciulli, P. W.. 2012. Revisiting dental fluctuating asymmetry in Neandertals and modern humans. American Journal of Physical Anthropology 149 (2): 193204.CrossRefGoogle ScholarPubMed
Barsky, D. 2013. The Caune de l’Arago stone industries in their stratigraphical context. Comptes Rendus Palevol 12 (5): 305–25.CrossRefGoogle Scholar
Barsky, D., Celiberti, V., Cauche, S. et al. 2010. Raw material discernment and technological aspects of the Barranco León and Fuente Nueva 3 stone assemblages (Orce southern Spain). Quaternary International 223–224: 201–19.Google Scholar
Barton, C. M. & Riel-Salvatore, J.. 2012. Agents of change: modeling biocultural evolution in Upper Pleistocene western Eurasia. Advances in Complex Systems 15 (1&2): 1150003.Google Scholar
Bar-Yosef, O. & Belfer-Cohen, A.. 2013. Following Pleistocene road signs of human dispersals across Eurasia. Quaternary International 285: 3043.Google Scholar
Bar-Yosef, O. & Bordes, J-G. 2010. Who were the makers of the Châtelperronian culture? Journal of Human Evolution 59: 586–93.CrossRefGoogle ScholarPubMed
Bataille, G. & Conard, N. J.. 2018. Blade and bladelet production at Hohle Fels Cave, AH IV in the Swabian Jura and its importance for characterizing the technological variability of the Aurignacian in Central Europe. PloS One 13 (4): e0194097.CrossRefGoogle ScholarPubMed
Becerra-Valdivia, L., Leal-Cervantes, R., Wood, R. & Higham, T.. 2020. Challenges in sample processing within radiocarbon dating and their impact in 14C dates-as-data studies. Journal of Archaeological Science 113: 105043.Google Scholar
Beck, M., Gaupp, R., Kamradt, I., Liebermann, C. & Pasda, C.. 2007. Bilzingsleben site formation processes-Geoarchaeological investigations of a Middle Pleistocene deposit: preliminary results of the 2003–2005 excavations. Archäologisches Korrespondenzblatt 37: 118.Google Scholar
Beier, J., Anthes, N., Wahl, J. & Harvati, K.. 2018. Similar cranial trauma prevalence among Neanderthals and Upper Palaeolithic modern humans. Nature 563: 686–90.Google Scholar
Beier, J., Anthes, N., Wahl, J. & Harvati, K.. 2020. Prevalence of cranial trauma in Eurasian Upper Paleolithic humans. American Journal of Physical Anthropology 174 (2): 268–84.Google Scholar
Bell, A. V., Hinde, K. & Newson, L.. 2013. Who was helping? The scope for female cooperative breeding in early Homo? PloS One 8 (12): e83667.CrossRefGoogle ScholarPubMed
Bello, S. M., Thomann, A., Signoli, M., Dutour, O. & Andrews, P.. 2006. Age and sex bias in the reconstruction of past population structures. American Journal of Physical Anthropology 129: 2438.Google Scholar
Bello, S. M., Saladié, P., Cáceres, I., Rodríguez-Hidalgo, A. & Parfitt, S. A.. 2015. Upper Palaeolithic ritualistic cannibalism at Gough’s Cave (Somerset, UK): the human remains from head to toe. Journal of Human Evolution 82: 170–89.Google Scholar
Benazzi, S., Douka, K., Fornai, C. et al. 2011. Early dispersal of modern humans in Europe and implications for Neanderthal behaviour. Nature 479: 525–9.Google Scholar
Benazzi, S., Slon, V., Talamo, S. et al. 2015. The makers of the Protoaurignacian and implications for Neandertal extinction. Science 348 (6236): 793–6.CrossRefGoogle ScholarPubMed
Benito, B. M., Svenning, J-C, Kellberg-Nielsen, T. et al. 2017. The ecological niche and distribution of Neanderthals during the Last Interglacial. Journal of Biogeography 44: 5161.Google Scholar
Benson, A., Kinsley, L., Willmes, M. et al. 2013. Laser ablation depth profiling of U-series and Sr isotopes in human fossils. Journal of Archaeological Science 40: 29913000.Google Scholar
Bentley, G. R. 1985. Hunter-gatherer energetics and fertility: a reassessment of the !Kung San. Human Ecology 13 (1): 79109.Google Scholar
Bentley, G. R., Goldberg, T. & Jasienska, G.. 1993a. The fertility of agricultural and non-agricultural traditional societies. Population Studies 47 (2): 269–81.Google Scholar
Bentley, G. R., Jasienska, G. & Goldberg, T.. 1993b. Is the fertility of agriculturalists higher than that of nonagriculturalists? Current Anthropology 34 (5): 778–85.CrossRefGoogle Scholar
Berger, T. D. & Trinkaus, E.. 1995. Patterns of trauma among the Neandertals. Journal of Archaeological Science 22 (6): 841–52.Google Scholar
Bergström, A., McCarthy, S. A., Hui, R. et al. 2020. Insights into human genetic variation and population history from 929 diverse genomes. Science 367 (6484). http://doi.org/10.1126/science.aay5012.CrossRefGoogle ScholarPubMed
Bermúdez de Castro, J. M. & Martinón-Torres, M. 2013. A new model for the evolution of the human Pleistocene populations of Europe. Quaternary International 295: 102–12.Google Scholar
Bermúdez de Castro, J. M. & Nicolás, M. E. 1997. Palaeodemography of the Atapuerca SH Middle Pleistocene hominid sample. Journal of Human Evolution 33: 333–55.Google Scholar
Bermúdez de Castro, J. M. & Rosas, A. 2001. Pattern of dental development in hominid XVII from the Middle Pleistocene Atapuerca Sima de los Huesos site (Spain). American Journal of Physical Anthropology 114: 325–30.Google Scholar
Bermúdez de Castro, J. M., Arsuaga, J. L, Carbonell, E, Rosas, A, Martínez, I & Mosquera, M. 1997. A hominid from the Lower Pleistocene of Atapuerca, Spain: possible ancestor to Neanderthals and modern humans. Science 276: 1392–5.CrossRefGoogle Scholar
Bermúdez de Castro, J. M., Martinón-Torres, M., Sarmiento, S., Arsuaga, J. L. & Carbonell, E.. 2003. Rates of anterior tooth wear in Middle Pleistocene hominins from Sima de los Huesos (Sierra de Atapuerca, Spain). Proceedings of the National Academy of Sciences USA 100 (21): 11992–6.Google Scholar
Bermúdez de Castro, J. M., Martinón-Torres, M., Lozano, M., Sarmiento, S. & Muela, A.. 2004. Paleodemography of the Atapuerca-Sima de los Huesos hominin sample: a revision and new approaches to the paleodemography of the European Middle Pleistocene population. Journal of Anthropological Research 60 (1): 526.Google Scholar
Bermúdez de Castro, J. M., Martinón-Torres, M, Prado, L. et al. 2010. New immature hominin fossil from European Lower Pleistocene shows the earliest evidence of a modern human developmental pattern. Proceedings of the National Academy of Sciences USA 107 (26): 11739–44.Google Scholar
Bermúdez de Castro, J. M., Martinón-Torres, M, Gómez-Robles, A. et al. 2011. Early Pleistocene human mandible from Sima del Elefante (TE) cave site in Sierra de Atapuerca (Spain): a comparative morphological study. Journal of Human Evolution 61: 1225.Google Scholar
Bermúdez de Castro, J. M., Martinón-Torres, M, Blasco, R., Rosell, J. & Carbonell, E.. 2013. Continuity or discontinuity in the European Early Pleistocene human settlement: the Atapuerca evidence. Quaternary Science Reviews 76: 5365.Google Scholar
Bermúdez de Castro, J. M., Martinón-Torres, M, Rosell, J., Blasco, R., Arsuaga, J. L. & Carbonell, E.. 2016. Continuity versus discontinuity of the human settlement of Europe between the late Early Pleistocene and the early Middle Pleistocene. The mandibular evidence. Quaternary Science Reviews 153: 5162.Google Scholar
Bermúdez de Castro, J. M., Martinón-Torres, M, Arsuaga, J. L. & Carbonell, E.. 2017. Twentieth anniversary of Homo antecessor (1997–2017): a review. Evolutionary Anthropology 26: 157–71.Google Scholar
Bertran, P., Sitzia, L., Banks, W. E. et al. 2013. The Landes de Gascogne (southwest France): periglacial desert and cultural frontier during the Palaeolithic. Journal of Archaeological Science 40 (5): 2274–85.Google Scholar
Bertranpretit, J. & Calafell, F.. 2002. From genetic variation to population dynamics: insights into the biological understanding of humans. In MacBeth, H. & Collinson, P. (eds.), Human Population Dynamics. Cross-Disciplinary Perspectives. Cambridge: Cambridge University Press, pp. 83102.Google Scholar
Bicho, N., Marrieros, J., Cascalheira, J., Pereira, T. & Haws, J.. 2015. Bayesian modelling and the chronology of the Portuguese Gravettian. Quaternary International 359–60: 499509.Google Scholar
Bicho, N., Cascalheira, J. & Gonçalves, C.. 2017. Early Upper Paleolithic colonization across Europe: time and mode of the Gravettian diffusion. PloS One 12 (5): e0178506.Google Scholar
Billari, F. C. 2015. Integrating macro-and micro-level approaches in the explanation of population change. Population Studies 69 (Supplement 1): S11S20.Google Scholar
Billy, G. 1975. Étude anthropologique des restes humains de l’Abri Pataud. Bulletin of the American School of Prehistoric Research 30: 201–61.Google Scholar
Binford, L. R. 2001. Constructing Frames of Reference: An Analytical Method for Archaeological Theory Building Using Ethnographic and Environmental Data Sets. Berkeley: University of California Press.Google Scholar
Binford, L. R. & Chasko, W. J. Jr. 1976. Nunamiut demographic history: a provocative case. In Zubrow, E. (ed.), Demographic Anthropology: Quantitative Approaches. Sante Fe: University of New Mexico Press, pp. 64144.Google Scholar
Binney, H., Edwards, M., Macias-Fauria, M. et al. 2017. Vegetation of Eurasia from the last glacial maximum to the present: key biogeographic patterns. Quaternary Science Reviews 157: 8097.Google Scholar
Biraben, J-N. 2003. The rising numbers of humankind. Population & Societies 394: 14.Google Scholar
Birch-Chapman, S., Jenkins, E., Coward, F. & Maltby, M.. 2017. Estimating population size, density and dynamics of pre-pottery Neolithic villages in the central and southern Levant: an analysis of Beidha, southern Jordan. Levant 49 (1): 123.Google Scholar
Bird, D. W. & O’Connell, J. F.. 2006. Behavioral ecology and archaeology. Journal of Archaeological Research 14 (2): 143–88.Google Scholar
Bird, D. W., Bliege Bird, R., Codding, B. F. & Zeanah, D. W.. 2019. Variability in the organization and size of hunter-gatherer groups: foragers do not live in small-scale societies. Journal of Human Evolution 131: 96108.Google Scholar
Bird, M. I., Ayliffe, L. K., Fifield, L. K. et al. 1999. Radiocarbon dating of ‘old’ charcoal using a wet oxidation, stepped-combustion procedure. Radiocarbon 41 (2): 127–40.CrossRefGoogle Scholar
Bird-David, N. 2018. Size matters! The scalability of modern hunter-gatherer animism. Quaternary International 464: 305–14.Google Scholar
Birdsell, J. B. 1953. Some environmental and cultural factors influencing the structuring of Australian Aboriginal populations. The American Naturalist 87 (834): 171207.Google Scholar
Birdsell, J. B. 1968. Some predictions for the Pleistocene based on equilibrium systems among recent hunter-gatherers. In Lee, R. B. & DeVore, I. (eds.), Man the Hunter. New York: Aldine, pp. 229–40.Google Scholar
Black, B. A., Neely, R. R. & Manga, M.. 2015. Campanian Igimbrite volcanism, climate, and the final decline of the Neanderthals. Geology 43 (5): 411–14.CrossRefGoogle Scholar
Blain, H. A., Cuenca-Bescós, G., Lozano-Fernández, I. et al. 2012. Investigating the Mid-Brunhes Event in the Spanish terrestrial sequence. Geology 40 (11): 1051–4.CrossRefGoogle Scholar
Bliege Bird, R. & Codding, B. F.. 2015. The sexual division of labor. In Scott, R & Kosslyn, S (eds.), Emerging Trends in the Social and Behavioral Sciences. Chichester: John Wiley & Son, pp. 116.Google Scholar
Blurton Jones, N. 1987. Bushman birth spacing: direct tests of some simple predictions. Ethology and Sociobiology 8: 183203.CrossRefGoogle Scholar
Blurton Jones, N. 2016. Demography and Evolutionary Ecology of Hadza Hunter-Gatherers. Cambridge: Cambridge University Press.Google Scholar
Blurton Jones., N. & Sibly, R.. 1978. Testing adaptiveness of culturally determined behaviour: do Bushman women maximise their reproductive success by spacing births widely and foraging seldom? In Blurton Jones, N. & Reynolds, V. (eds.), Human Behaviour and Adaptation: London: Taylor and Francis, pp. 135–55.Google Scholar
Blurton Jones, N. G., Hawkes, K & O’Connell, J. F.. 2002. Antiquity of postreproductive life: are there modern impacts on hunter-gatherer postreproductive lifespans? American Journal of Human Biology 14: 184205.Google Scholar
Bocherens, H., Billiou, D., Mariotti, A. et al. 2001. New isotopic evidence for dietary habits of Neanderthals from Belgium. Journal of Human Evolution 40: 497505.Google Scholar
Bocquet-Appel, J.-P. 2008. Recent Advances in Paleodemography: Data, Techniques, Patterns. Dordrecht: Springer.Google Scholar
Bocquet-Appel, J.-P. 2011. When the world’s population took off: the springboard of the Neolithic Demographic Transition. Science 333 (6042): 560–1.Google Scholar
Bocquet-Appel, J.-P. & Arsuaga, J.-L.. 1999. Age distribution of hominid samples at Atapuerca (SH) and Krapina could indicate accumulation by catastrophe. Journal of Archaeological Science 26: 327–38.Google Scholar
Bocquet-Appel, J.-P. & Bar-Yosef, O.. 2008. The Neolithic Demographic Transition and Its Consequences. Dordrecht: Springer.Google Scholar
Bocquet-Appel, J.-P. & Demars, P.- Y.. 2000. Population kinetics in the Upper Palaeolithic in western Europe. Journal of Archaeological Science 27 (7): 551–70.Google Scholar
Bocquet-Appel, J.-P. & Degioanni, A.. 2013. Neanderthal demographic estimates. Current Anthropology 54 (S8): S202–13.CrossRefGoogle Scholar
Bocquet-Appel, J.-P. & Masset, C.. 1982. Farewell to Paleodemography. Journal of Human Evolution 11: 321–33.Google Scholar
Bocquet-Appel, J.-P. & Tuffreau, A.. 2009. Technological responses to macroclimatic variations (240,000–40,000 BP). Human Biology 81 (2/3): 287307.Google Scholar
Bocquet-Appel, J.-P., Demars, P.-Y., Noiret, L. & Dobrowsky, D.. 2005. Estimates of Upper Palaeolithic meta-populations size in Europe from archaeological data. Journal of Archaeological Science 32: 1656–68.Google Scholar
Bogin, B. & Rios, L.. 2003. Rapid morphological change in living humans: implications for modern human origins. Comparative Biochemistry and Physiology Part A 136: 7184.Google Scholar
Bon, F. & Bodu, P.. 2002. Analyse technologique du débitage aurignacien. In Schmider, B. (ed.), L’Aurignacien de la Grotte du Renne: Les fouilles d’André Leroi-Gourhan à Arcy sur-Cure (Yonne). XXXIVe supplément à Gallia Préhistoire. Paris: CNRS Editions, pp. 115133.Google Scholar
Bonjean, D., Vanbrabant, Y., Abrams, G. et al. 2015. A new Cambrian black pigment used during the late Middle Palaeolithic discovered at Scladina Cave (Andenne, Belgium). Journal of Archaeological Science 55: 253–65.Google Scholar
Bonmatí, A., Gómez-Olivencia, A., Arsuaga, J-L et al. 2010. Middle Pleistocene lower back and pelvis from an aged human individual from the Sima de los Huesos site, Spain. Proceedings of the National Academy of Sciences USA 107 (43): 18386–91.Google Scholar
Boone, J. L. 2002. Subsistence strategies and early human population history: an evolutionary ecological perspective. World Archaeology 31: 625.CrossRefGoogle Scholar
Bordes, J.-G. & Teyssandier, N.. 2011. The Upper Palaeolithic nature of the Châtelperronian in south-western France: archeostratigraphic and lithic evidence. Quaternary International 246 (1–2): 382–8.Google Scholar
Borgerhoff Mulder, M. & Schacht, R.. 2012. Human Behavioural Ecology, in eLS. Chichester: John Wiley & Sons. http://doi.org/10.1002/9780470015902.a.0003671.pub2.Google Scholar
Borgia, V. 2017. Hunting high and low: Gravettian hunting weapons from southern Italy to the Russian Plain. Open Archaeology 3: 376–91.Google Scholar
Borić, D. & Cristiani, E. 2016. Social networks and connectivity among the Palaeolithic and Mesolithic foragers of the Balkans and Italy. In Raiko, K & Floss, H (eds.), Southeast Europe before Neolithisation (Proceedings of the Workshop Held in Tübingen, Germany, 9 May 2014). Tübingen: Pro BUSINESS digital printing Deutschland GmbH, pp. 73112.Google Scholar
Boschin, F. 2019. Exploitation of carnivores, lagomorphs and rodents at Grotta Paglicci during the Epigravettian: the dawn of a new subsistence strategy. Journal of Archaeological Science: Reports 26: 101871.Google Scholar
Boserup, E. 1965 [2003]. The Conditions of Agricultural Growth: The Economics of Agrarian Change Under Population Pressure. London: Routledge.Google Scholar
Bosinski, G. 1966. Der paläolithische Fundplatz Rheindahlen, Ziegelei Dreesen-Westwand. Bonner Jahrbücher 166: 318–48.Google Scholar
Bosinski, G., Brunnacker, K. & Turner, E.. 1983. Ein Siedlungsbefund des frühen Mittelpaläolithikums von Ariendorf, KR. Neuwied. Archäologisches Korrespondenzblatt 13: 157–69.Google Scholar
Bourguignon, L., Sellami, F., Deloze, V., Sellier-Segard, N., Beyries, S. & Emery-Barbier, A.. 2002. L’habitat moustérien de La Folie (Poitiers, Vienne): synthèse des premiers resultants. Paléo 14: 2948.Google Scholar
Bourguignon, L., Crochet, J-Y, Capdevila, R. et al. 2016. Bois de Riquet (Lézignan-la-Cèbe, Hérault): a late Early Pleistocene archaeological occurrence in southern France. Quaternary International 393: 2440.Google Scholar
Bourrillon, R., White, R., Tartar, E. et al. 2018. A new Aurignacian engraving from Abri Blanchard, France: implications for understanding Aurignacian graphic expression in Western and Central Europe. Quaternary International 491: 4664.CrossRefGoogle Scholar
Boyd, R. & Richerson, P. J.. 1985. Culture and the Evolutionary Process. Chicago: University of Chicago Press.Google Scholar
Bradshaw, C. J. A., Cooper, A., Turney, C. S. M. & Brook, B. W.. 2012. Robust estimates of extinction time in the geological record. Quaternary Science Reviews 33: 1419.Google Scholar
Bradshaw, C. J. A., Ulm, S., Williams, A. N. et al. 2019. Minimum founding populations for the first peopling of Sahul. Nature Ecology & Evolution 3 (7): 1057–63.Google Scholar
Bradtmöller, M., Pastoors, A., Weninger, B. & Weniger, G.-C.. 2012. The repeated replacement model–rapid climate change and population dynamics in Late Pleistocene Europe. Quaternary International 247: 3849.Google Scholar
Brewster, C., Meiklejohn, C., von Cramon-Taubadel, N. & Pinhasi, R.. 2014. Craniometric analysis of European Upper Palaeolithic and Mesolithic samples supports discontinuity at the Last Glacial Maximum. Nature Communications 5 (1): 110.Google Scholar
Briggs, A. W., Good, J. M., Green, R. E. et al. 2009. Targeted retrieval and analysis of five Neanderthal mtDNA genomes. Science 325: 318–21.Google Scholar
Bromham, L., Hua, X., Fitzpatrick, T. G. & Greenhill, S. J.. 2015. Rate of language evolution is affected by population size. Proceedings of the National Academy of Sciences USA 112 (7): 2097–102.Google Scholar
Brown, W. A. 2015. Through a filter, darkly: population size estimation, systematic error, and random error in radiocarbon-supported demographic temporal frequency analysis. Journal of Archaeological Science 53: 133–47.Google Scholar
Buck, L. T. & Stringer, C. B.. 2014. Having the stomach for it: a contribution to Neanderthal diets? Quaternary Science Reviews 96: 161–67.Google Scholar
Burger, O., Baudisch, A. & Vaupel, J. W.. 2012. Human mortality improvement in evolutionary context. Proceedings of the National Academy of Sciences USA 109 (44): 18210–14.Google Scholar
Burke, A. 2006. Neanderthal settlement patterns in Crimea: a landscape approach. Journal of Anthropological Archaeology 25: 510–23.Google Scholar
Burke, A. 2012. Spatial abilities, cognition and the pattern of Neanderthal and modern human dispersals. Quaternary International 247: 230–35.Google Scholar
Burke, A., Levavasseur, G., James, P. M. A. et al. 2014. Exploring the impact of climate variability during the Last Glacial Maximum on the pattern of human occupation of Iberia. Journal of Human Evolution 73: 3546.Google Scholar
Burke, A., Kageyama, M., Latombe, G. et al. 2017. Risky Business: the impact of climate and climate variability on human population dynamics in Western Europe during the Last Glacial Maximum. Quaternary Science Reviews 164: 217–29.Google Scholar
Burke, A., Riel-Salvatore, J. & Barton, C. M.. 2018. Human response to habitat suitability during the Last Glacial Maximum in Western Europe. Journal of Quaternary Science 33 (3): 335–45.Google Scholar
Caldwell, J. & Caldwell, P.. 1983. The demographic evidence for the incidence and cause of abnormally low fertility in tropical Africa. World Health Statistics Quarterly 36: 234.Google ScholarPubMed
Camarós, E., Cueto, M., Lorenzo, C., Villaverde, V. & Rivals, F.. 2016. Large carnivore attacks on hominins during the Pleistocene: a forensic approach with a Neanderthal example. Archaeological and Anthropological Sciences 8: 635–46.Google Scholar
Camarós, E., Cueto, M., Rosell, J. et al. 2017. Hunted or scavenged Neanderthals? Taphonomic approach to hominin fossils with carnivore damage. International Journal of Osteoarchaeology 27: 606–20.Google Scholar
Campbell, K. L. & Wood, J. W.. 1988. Fertility in traditional societies. In Diggory, P., Potts, M. & Teper, S. (eds.), Natural Human Fertility. London: Palgrave MacMillan, pp. 3969.CrossRefGoogle Scholar
Candy, I. & McClymont, E. L.. 2013. Interglacial intensity in the North Atlantic over the last 800 000 years: investigating the complexity of the mid‐Brunhes Event. Journal of Quaternary Science 28 (4): 343–8.Google Scholar
Carbonell, E. & Mosquera, M.. 2006. The emergence of a symbolic behaviour: the sepulchral pit of Sima de los Huesos, Sierra de Atapuerca, Burgos, Spain. Comptes Rendus Palevol 5 (1–2): 155–60.Google Scholar
Carbonell, E., Mosquera, M., Rodrıguez, X. P., Sala, R. & van der Made, J.. 1999. Out of Africa: the dispersal of the earliest technical systems reconsidered. Journal of Anthropological Archaeology 18 (2): 119–36.Google Scholar
Carbonell, E., Bermúdez de Castro, J. M, Parés, J. M et al. 2008. The first hominin of Europe. Nature 452: 465–70.Google Scholar
Carbonell, E., Cácares, I., Lozano, M. et al. 2010. Cultural cannibalism as a paleoeconomic system in the European Lower Pleistocene. The case of level TD6 of Gran Dolina (Sierra de Atapuerca, Burgos, Spain). Current Anthropology 51 (4): 539–49.CrossRefGoogle Scholar
Caron, F., d’Errico, F, Del Moral, P, Santos, F & Zilhão, J. 2011. The reality of Neanderthal symbolic behaviour at the Grotte du Renne, Arcy-sur-Cure, France. PloS One 6 (6): e21545.Google Scholar
Carotenuto, F., Tsikaridze, N., Rook, L. et al. 2016. Venturing out safely: the biogeography of Homo erectus dispersal out of Africa. Journal of Human Evolution 95: 112.Google Scholar
Carretero, J-M., Rodríguez, L., Garćia-González, R. et al. 2012. Stature estimation from complete long bones in the Middle Pleistocene humans from the Sima de los Huesos, Sierra de Atapuerca (Spain). Journal of Human Evolution 62: 242–55.Google Scholar
Carter, T., Contreras, D. A., Holcomb, J. et al. 2019. Earliest occupation of the Central Aegean (Naxos), Greece: Implications for hominin and Homo sapiens’ behavior and dispersals. Science Advances 5 (10): p.eaax0997.Google Scholar
Caspari, R. & Lee, S.-H.. 2004. Older age becomes common late in human evolution. Proceedings of the National Academy of Sciences USA 101 (30): 10895–900.Google Scholar
Caspari, R & Lee, S.-H.. 2006. Is human longevity a consequence of cultural change or modern biology? American Journal of Physical Anthropology 129: 512–17.Google Scholar
Caspari, R., Rosenberg, K. R. & Wolpoff, M. H.. 2017. Brother or other: the place of Neanderthals in human evolution. In Marom, A & Hovers, E (eds.), Human Paleontology and Prehistory: Contributions in Honor of Yoel Rak. Dordrecht: Springer, pp. 253–71.Google Scholar
Castellano, S., Parra, G., Sánchez-Quinto, F. A. et al. 2014. Patterns of coding variation in the complete exomes of three Neanderthals. Proceedings of the National Academy of Sciences USA 111 (18): 6666–71.Google Scholar
Cattin, M-I. 2018. Moving with the Magdalenians: examples from the camp sites of Monruz and Champréveyres (Switzerland). Quaternary International 498: 411.Google Scholar
Cavalli-Sforza, L. L. 1986. Demographic data. In Cavalli-Sforza, L. L (ed.), African Pygmies: New York: Academic Press, pp. 2344.Google Scholar
Charnov, E. L & Berrigan, D.. 1993. Why do female primates have such long lifespans and so few babies? or life in the slow lane. Evolutionary Anthropology 1: 191–4.Google Scholar
Chen, F., Welker, F., Shen, C. C. et al. 2019. A late Middle Pleistocene Denisovan mandible from the Tibetan plateau. Nature 569 (7756): 409–12.Google Scholar
Chen, L., Wolf, A. B., Fu, W., Li, L. & Akey, J. M. 2020. Identifying and interpreting apparent Neanderthal ancestry in African individuals. Cell 180 (4): 677–87.Google Scholar
Chevalier, T. 2019. Trauma in the upper limb of an Upper Paleolithic female from Caviglione cave (Liguria, Italy): Etiology and after-effects in bone biomechanical properties. International Journal of Paleopathology 24: 94107.Google Scholar
Chiotti, L., Nespoulet, R. & Henry-Gambier, D.. 2015. Occupations and status of the Abri Pataud (Dordogne, France) during the Final Gravettian. Quaternary International 359–60: 406–22.Google Scholar
Chu, W. 2018. The Danube Corridor hypothesis and the Carpathian Basin: geological, environmental and archaeological approaches to characterizing Aurignacian dynamics. Journal of World Prehistory 31 (2): 117–78.Google Scholar
Churchill, S. E. 2014. Thin on the Ground: Neandertal Biology, Archeology, and Ecology. Oxford: Wiley Blackwell.Google Scholar
Churchill, S. E. & Smith, F. H.. 2000. Makers of the early Aurignacian of Europe. American Journal of Physical Anthropology 113 (S31): 61115.Google Scholar
Churchill, S. E., Formicola, V., Holliday, T. W., Holt, B. M. & Schumann, B. A.. 2000. The Upper Palaeolithic population of Europe in an evolutionary perspective. In Roebroeks, W., Mussi, M., Svoboda, J. & Fennema, K. (eds.), Hunters of the Golden Age: The Mid Upper Palaeolithic of Eurasia 30,000–20,000 BP. Leiden: University of Leiden Press, pp. 3157.Google Scholar
Churchill, S. E., Franciscus, R. G., McKean-Peraza, H. A., Daniel, J. A. & Warren, B. R.. 2009. Shanidar 3 Neanderthal rib puncture wound and Palaeolithic weaponry. Journal of Human Evolution 57: 163–78.Google Scholar
Churchill, S. E., Walker, C. S. & Schwartz, A. M.. 2016. Home-range size in large-bodied carnivores as a model for predicting Neandertal territory size. Evolutionary Anthropology 25: 117–23.Google Scholar
Clark, G. 1969. World Prehistory: A New Synthesis. Cambridge: Cambridge University Press.Google Scholar
Clark, G. A., Barton, C. M. & Straus, L. G.. 2019. Landscapes, climate change & forager mobility in the Upper Paleolithic of northern Spain. Quaternary International 515: 176–87.Google Scholar
Clark, P. U., Dyke, A. S., Shakun, J. D. et al. 2009. The last Glacial Maximum. Science 325 (5941): 710–14.Google Scholar
Clarkson, C., Jacobs, Z., Marwick, B. et al. 2017. Human occupation of northern Australia by 65,000 years ago. Nature 547: 306–13.Google Scholar
Codding, B. F. & Bird, D. W.. 2015. Behavioral ecology and the future of archaeological science. Journal of Archaeological Science 56: 920.Google Scholar
Cohen, K. M., MacDonald, K., Joordens, J . C. A., Roebroeks, W. & Gibbard, P. L.. 2012. The earliest occupation of north-west Europe: a coastal perspective. Quaternary International 271: 7083.Google Scholar
Colchero, F., Rau, R., Jones, O. R. et al. 2016. The emergence of longevous populations. Proceedings of the National Academy of Sciences USA 113 (48): E7681–90.Google Scholar
Cole, J. 2017. Assessing the calorific significance of episodes of human cannibalism in the Palaeolithic. Scientific Reports 7: 44707.Google Scholar
Collard, M., Buchanan, B., Morin, J. & Costopoulos, A.. 2011. What drives the evolution of hunter-gatherer subsistence technology? A reanalysis of the risk hypothesis with data from the Pacific Northwest. Philosophical Transactions of the Royal Society B 366: 1129–38.Google Scholar
Collard, M., Buchanan, B. & O’Brien, M. J.. 2013a. Population size as an explanation for patterns in the Paleolithic archaeological record: more caution is needed. Current Anthropology 54 (S8): S388S396.Google Scholar
Collard, M., Buchanan, B., O’Brien, M. J. & Scholnick, J.. 2013b. Risk, mobility or population size? Drivers of technological richness among contact-period western North American hunter-gatherers. Philosophical Transactions of the Royal Society B 368: 2012412.Google Scholar
Collard, M., Vaesen, K., Cosgrove, R. & Roebroeks, W.. 2016. The empirical case against the ‘demographic turn’ in Palaeolithic archaeology. Philosophical Transactions of the Royal Society B 371: 20150242.Google Scholar
Conard, N. J. & Bolus, M.. 2003. Radiocarbon dating the appearance of modern humans and timing of cultural innovations in Europe: new results and new challenges. Journal of Human Evolution 44 (3): 331–71.Google Scholar
Conard, N. J., Grootes, P. M. & Smith, F. H.. 2004. Unexpectedly recent dates for human remains from Vogelherd. Nature 430 (6996): 198201.Google Scholar
Conard, N. J., Bolus, M. & Münzel, S. C.. 2012. Middle Paleolithic land use, spatial organization and settlement intensity in the Swabian Jura, southwestern Germany. Quaternary International 247: 236245.Google Scholar
Conard, N. J., Serangeli, J., Böhner, U. et al. 2015. Excavations at Schöningen and paradigm shifts in human evolution. Journal of Human Evolution 89: 117.Google Scholar
Conkey, M. W. 1980. The identification of prehistoric aggregation sites: the case of Altamira. Current Anthropology 21 (5): 609–30.Google Scholar
Conkey, M. W. & Spector, J. D.. 1984. Archaeology and the study of gender. Advances in Archaeological Method and Theory 7: 138.Google Scholar
Contreras, D. A. & Meadows, J.. 2014. Summed radiocarbon calibrations as a population proxy: a critical evaluation using a realistic simulation approach. Journal of Archaeological Science 52: 591608.Google Scholar
Cook, J. 2015. Was bedeutet ein Name? Ein Rückblick auf die Ursprünge, Geschichte und Unangemessenheit des Begriffs Venusfigur. Die Kunde: Zeitschrift für Niedersächsische Archäologie 66: 4372.Google Scholar
Corbey, R., Jagich, A., Vaesen, K. & Collard, M.. 2016. The Acheulean handaxe: More like a bird’s song than a beatles’ tune? Evolutionary Anthropology 25 (1): 619.Google Scholar
Cortés-Sánchez, M., Jiménez-Espejo, F. J., Simón-Vallejo, M. D. et al. 2019. An early Aurignacian arrival in southwestern Europe. Nature Ecology and Evolution 3 (2): 207–12.Google Scholar
Coward, F. 2016. Scaling up: material culture as scaffold for the social brain. Quaternary International 405: 7890.Google Scholar
Coward, F. & Gamble, C.. 2008. Big brains, small worlds: material culture and human evolution. Philosophical Transactions of the Royal Society B 363: 1969–79.Google Scholar
Cowgill, L. W., Mednikova, M. B., Buzhilova, A. P. & Trinkaus, E.. 2015. The Sunghir 3 Upper Palaeolithic juvenile: pathology versus persistence in the Palaeolithic. International Journal of Osteoarchaeology 25: 176–87.Google Scholar
Cox, S. L., Ruff, C. B., Maier, R. M. & Mathieson, I.. 2019. Genetic contributions to variation in human stature in prehistoric Europe. Proceedings of the National Academy of Sciences USA 116 (43): 21484–92.Google Scholar
Craig, O. E., Biazzo, M., Colonese, A. C. et al. 2010. Stable isotope analysis of Late Upper Palaeolithic human and faunal remains from Grotta del Romito (Cosenza, Italy). Journal of Archaeological Science 37: 2504–12.Google Scholar
Creanza, N. & Feldman, M. W.. 2016. Worldwide genetic and cultural change in human evolution. Current Opinion in Genetics and Development 41: 8592.Google Scholar
Creanza, N., Kolodny, O. & Feldman, M. W.. 2017a. Cultural evolutionary theory: how culture evolves and why it matters. Proceedings of the National Academy of Sciences USA 114 (30): 7782–9.Google Scholar
Creanza, N., Kolodny, O. & Feldman, M. W.. 2017b. Greater than the sum of its parts? Modelling population contact and interaction of cultural repertoires. Journal of the Royal Society Interface 14: 20170171.Google Scholar
Crema, E. R., Bevan, A. & Shennan, S.. 2017. Spatio-temporal approaches to archaeological radiocarbon dates. Journal of Archaeological Science 87: 19.Google Scholar
Crevecoeur, I., Bayle, P., Rougier, H. et al. 2010. The Spy VI child: a newly discovered Neanderthal infant. Journal of Human Evolution 59: 641–56.Google Scholar
Crochet, J-Y., Welcomme, J-L, Ivorra, J. et al. 2009. Une nouvelle faune de vertébrés continentaux associée à des artefacts dans le Pléistocène inférieur de l’Hérault (Sud de la France), vers 1,57 Ma. Comptes Rendus Palevol 8: 725736.Google Scholar
Crubézy, E. & Trinkaus, E.. 1992. Shanidar 1: a case of hyperostotic disease (DISH) in the Middle Paleolithic. American Journal of Physical Anthropology 89: 411420.Google Scholar
Cucart-Mora, C., Lozano, S. & de Pablo, J. F. L.. 2018. Bio-cultural interactions and demography during the Middle to Upper Palaeolithic transition in Iberia: an agent-based modelling approach. Journal of Archaeological Science 89: 1424.Google Scholar
Cunha, E., Ramirez Rozzi, F., Bermúdez de Castro, J. M, Martinón-Torres, M, Wasterlain, S. N & Sarmiento, S. 2004. Enamel hypoplasias and physiological stress in the Sima de los Huesos Middle Pleistocene hominins. American Journal of Physical Anthropology 125: 220–31.Google Scholar
Cunningham, A. J., Worthington, S., Venkataraman, V. V. & Wrangham, R. W.. 2019. Do modern hunter-gatherers live in marginal environments? Journal of Archaeological Science: Reports 25: 584–99.Google Scholar
d’Errico, F. 2003. The invisible frontier: a multiple species model for the origin of behavioral modernity. Evolutionary Anthropology 12 (4): 188202.Google Scholar
d’Errico, F. & Vanhaeren, M.. 2015. Upper Palaeolithic mortuary practices: reflection of ethnic affiliation, social complexity, and cultural turnover. In Renfrew, C, Boyd, M & Morley, I (eds.), Death Rituals, Social Order and the Archaeology of Immortality in the Ancient World: ‘Death shall have no dominion’. Cambridge: Cambridge University Press, pp. 4561.Google Scholar
Dalén, L., Orlando, L., Shapiro, B. et al. 2012. Partial genetic turnover in Neandertals: continuity in the east and population replacement in the west. Molecular Biology and Evolution 29 (8): 18931897.Google Scholar
Daura, J., Sanz, M., Arsuaga, J. L. et al. 2017. New Middle Pleistocene hominin cranium from Gruta da Aroeira (Portugal). Proceedings of the National Academy of Sciences USA 114 (3): 33973402.Google Scholar
David-Barrett, T. 2019. Network effects of demographic transition. Scientific Reports 9: 2361.Google Scholar
Davies, W. 2001. A very model of a modern human industry: new perspectives on the origins and spread of the Aurignacian in Europe. Proceedings of the Prehistoric Society 67: 195217.Google Scholar
Davies, W. & Gollop, P.. 2003. The human presence in Europe during the Last Glacial period II: Climate tolerances and climate preferences of mid-and late Glacial hominids. In van Andel, T. H. & Davies, W. (eds.), Neanderthals and Modern Humans in the European Landscape during the Last Glaciation: Archaeological Results of the Stage 3 Project. Cambridge: McDonald Institute Monographs, pp. 131–46.Google Scholar
Davies, W., White, D., Lewis, M. & Stringer, C.. 2015. Evaluating the transitional mosaic: frameworks of change from Neanderthals to Homo sapiens in eastern Europe. Quaternary Science Reviews 118: 211–42.Google Scholar
Davis, R. & Ashton, N.. 2019. Landscapes, environments and societies: The development of culture in Lower Palaeolithic Europe. Journal of Anthropological Archaeology 56: 101107.Google Scholar
De Fanti, S., Barbieri, C., Sarno, S. et al. 2015. Fine dissection of human mitochondrial DNA haplogroup HV lineages reveal Palaeolithic signatures from European Glacial Refugia. Plos One 10 (12): e0144391.Google Scholar
de la Peña, P. & Toscano, G. V.. 2013. The Early Upper Palaeolithic puzzle in Mediterranean Iberia. Quartär 60: 85106.Google Scholar
de Lumley, H. 1969. A Paleolithic camp at Nice. Scientific American 220: 4250.Google Scholar
Dean, M. C. 2016. Measures of maturation in early fossil hominins: events at the first transition from australopiths to early Homo. Philosophical Transactions of the Royal Society B 371: 20150234.Google Scholar
Dediu, D. & Levinson, S. C.. 2018. Neanderthal language revisited: not only us. Current Opinion in Behavioral Sciences 21: 4955.Google Scholar
Defleur, A., White, T., Valensi, P., Slimak, L. & Crégut-Bonnoure, E.. 1999. Neanderthal cannibalism at Moula-Guercy, Ardèche, France. Science 286: 128–31.Google Scholar
Defleur, A. R. & Desclaux, E.. 2019. Impact of the last interglacial climate change on ecosystems and Neanderthals behaviour at Baume Moula-Guercy, Ardèche, France. Journal of Archaeological Science 104: 114–24.Google Scholar
Defleur, A. R., Desclaux, E., Jabbour, R. S. & Richards, G. D.. 2020. The Eemian: global warming, ecosystem upheaval, demographic collapse and cannibalism at Moula-Guercy: a reply to Slimak and Nicholson (2020). Journal of Archaeological Science 117: 105113.Google Scholar
Degioanni, A., Bonnefant, C., Cabut, S. & Condemi, S.. 2019. Living on the edge: was demographic weakness the cause of Neanderthal demise? PloS One 14 (5): e0216742.Google Scholar
Delpech, F. 1999. Biomasse d’Ongulés au Paléolithique et inférences sur la démographie. Paléo, 11 (1): 1942.Google Scholar
Delporte, H. 1993. Gravettian female figurines: a regional survey. In Knecht, H., Pike-Tay, A. & White, R. (eds.), Before Lascaux: The Complex Record of the Early Upper Paleolithic. Boca Raton: CRC Press, pp. 243–25.Google Scholar
Demars, P.-Y. 1996. Démographie et occupation de l’espace au Paléolithique supérieur et au Mésolithique en France. Préhistoire Européenne 8: 326.Google Scholar
Demidenko, Y. E. In Press. South of Eastern Europe and Upper Paleolithic diversity around the Last Glacial Maximum. Quaternary International (2020): https://doi.org/10.1016/j.quaint.2020.07.002.Google Scholar
Demuro, M., Arnold, L. J., Duval, M., Méndez-Quintas, E., Santonja, M. & Pérez-González, A.. 2020. Refining the chronology of Acheulean deposits at Porto Maior in the River Miño basin (Galicia, Spain) using a comparative luminescence and ESR dating approach. Quaternary International 556: 96112.Google Scholar
Dennell, R. 2009. The Palaeolithic Settlement of Asia. Cambridge: Cambridge University Press.Google Scholar
Dennell, R. 2017. Pleistocene hominin dispersals, naïve faunas and social networks. In Boivin, N., Crassard, R. & Petraglia, M. (eds.), Human Dispersal and Species Movement from Prehistory to the Present. Cambridge: Cambridge University Press, pp. 6289.Google Scholar
Dennell, R. & Roebroeks, W.. 1996. The earliest colonization of Europe: the short chronology revisited. Antiquity 70 (269): 535–42.Google Scholar
Dennell, R. W., Martinón-Torres, M & Bermúdez de Castro, J. M. 2011. Hominin variability, climatic instability and population demography in Middle Pleistocene Europe. Quaternary Science Reviews 30 (11–12): 1511–24.Google Scholar
Dennell, R., Martinón-Torres, M., Bermúdez de Castro, J. M & Xing, G. 2020. A demographic history of late Pleistocene China. Quaternary International 559: 413.Google Scholar
Derex, M., Beugin, M.-P., Godelle, B. & Raymond, M.. 2013. Experimental evidence for the influence of group size on cultural complexity. Nature 503: 389–91.Google Scholar
Derex, M., Perreault, C. & Boyd, R.. 2018. Divide and conquer: intermediate levels of population fragmentation maximize cultural accumulation. Philosophical Transactions of the Royal Society B 373: 20170062.Google Scholar
Despriée, J., Voinchet, P., Tissoux, H. et al. 2011. Lower and Middle Pleistocene human settlements recorded in fluvial deposits of the middle Loire River Basin, Centre region, France. Quaternary Science Reviews 30: 1474–85.Google Scholar
Deviese, T., Comeskey, D., McCullagh, J., Bronk Ramsey, C. & Higham, T.. 2018. New protocol for compound‐specific radiocarbon analysis of archaeological bones. Rapid Communications in Mass Spectrometry 32 (5): 373–9.Google Scholar
Devièse, T., Abrams, G., Hajdinjak, M. et al. 2021. Reevaluating the timing of Neanderthal disappearance in Northwest Europe. Proceedings of the National Academy of Sciences 118 (12): e2022466118.Google Scholar
Di Modica, K., Toussaint, M., Abrams, G. & Pirson, S.. 2016. The Middle Palaeolithic from Belgium: Chronostratigraphy, territorial management and culture on a mosaic on contrasting environments. Quaternary International 411: 77106.Google Scholar
Dibble, H. L., Sandgathe, D., Goldberg, P., McPherron, S. P & Aldeias, V.. 2018. Were Western European Neandertals able to make fire? Journal of Paleolithic Archaeology 1 (1): 5479.Google Scholar
Dickeman, M. 1975. Demographic consequences of infanticide in man. Annual Review of Ecology and Systematics 6 (1): 107–37.Google Scholar
Diekmann, Y., Smith, D., Gerbault, P. et al. 2017. Accurate age estimation in small-scale societies. Proceedings of the National Academy of Sciences USA 114 (31): 8205–10.Google Scholar
Dinnis, R. 2015. A survey of northwestern European Aurignacian sites and some comments regarding their potential chrono-cultural significance. In Ashton, N. & Harris, C. (eds.), No Stone Unturned: Papers in Honour of Roger Jacobi. Lithic Studies Society Occasional Paper 9. London: Lithic Studies Society, pp. 5976.Google Scholar
Dinnis, R., Bessudnov, A., Reynolds, N. et al. 2019a. New data for the Early Upper Palaeolithic of Kostenki (Russia). Journal of Human Evolution 127: 2140.Google Scholar
Dinnis, R., Bessudnov, A., Chiotti, L., Flas, D. & Michel, A.. 2019b. Thoughts on the structure of the European Aurignacian, with particular focus on Hohle Fels IV. Proceedings of the Prehistoric Society 85: 2960.Google Scholar
Divale, W. T. 1972. Systemic population control in the Middle and Upper Palaeolithic: inferences based on contemporary hunter-gatherers. World Archaeology 4 (2): 222243.CrossRefGoogle ScholarPubMed
Djindjian, F. 2015. Identifying the hunter-gatherer systems behind associated mammoth bone beds and mammoth bone dwellings. Quaternary International 359–60: 4757.Google Scholar
Dobzhansky, T. 1935. A critique of the species concept in biology. Philosophy of Science 2: 344–55.Google Scholar
Dogandžić, T. & McPherron, S. P.. 2013. Demography and the demise of Neandertals: a comment on ‘Tenfold Population Increase at the Neandertal-to-Modern-Human Transition’. Journal of Human Evolution 64 (4): 311–13.Google Scholar
Dogandžić, T., McPherron, S & Mihailović, D. 2014. Middle and Upper Palaeolithic in the Balkans: continuities and discontinuities of human occupations. In Milhailović, D. (ed.), Palaeolithic and Mesolithic Research in the Central Balkans. Belgrade: Serbian Archaeological Society, pp. 8396.Google Scholar
Dolukhanov, P., Sokoloff, S. & Shukurov, A.. 2001. Radiocarbon chronology of Upper Palaeolithic sites in Eastern Europe at improved resolution. Journal of Archaeological Science 28: 699712.Google Scholar
Doronichev, V. 2008. The Lower Palaeolithic in Eastern Europe and the Caucasus: a reappraisal of the data and new approaches. PaleoAnthropology 2008: 107–57.Google Scholar
Doronichev, V. 2016. The pre-Mousterian industrial complex in Europe between 400 and 300 ka: interpreting its origin and spatiotemporal variability. Quaternary International 409: 222–40.Google Scholar
Doronichev, V. & Golovanova, L.. 2010. Beyond the Acheulean: a view on the Lower Palaeolithic occupation of Western Eurasia. Quaternary International 223–34: 327–44.Google Scholar
Doronicheva, E. V., Kulkova, M. A. & Shackley, M. S.. 2014. Raw material exploitation, transport, and mobility in the Northern Caucasus Eastern Micoquian. PaleoAnthropology 2016: 145.Google Scholar
Douka, K., Higham, T. F. G., Wood, R. et al. 2014. On the chronology of the Uluzzian. Journal of Human Evolution 68: 113.Google Scholar
Douka, K., Slon, V., Jacobs, Z. et al. 2019. Age estimates for hominin fossils and the onset of the Upper Palaeolithic at Denisova Cave. Nature 565: 640–5.Google Scholar
Drennan, R. D, Berrey, C. A. & Peterson, C. E.. 2015. Regional Settlement Demography in Archaeology. Clinton Corners, NY: Eliot Werner Publications Inc.Google Scholar
Duarte, C., Maurício, J., Pettitt, P. B. et al. 1999. The early Upper Paleolithic human skeleton from the Abrigo do Lagar Velho (Portugal) and modern human emergence in Iberia. Proceedings of the National Academy of Sciences USA 96 (13): 7604–9.Google Scholar
Ducasse, S. 2012. What is left of the Badegoulian ‘interlude’? New data on cultural evolution in southern France between 23,500 and 20,500 cal BP. Quaternary International 272–3: 150–65.Google Scholar
Ducasse, S., Pétillon, J-M, Chauvière, F-X, Renard, C., Lacrampe-Cuyaubère, F. & Muth, X.. 2019. Archaeological recontextualization and first direct 14C dating of a ‘pseudo-excise’ decorated antler point from France (Pégourié Cave, Lot): implications on the cultural geography of southwestern Europe during the Last Glacial Maximum. Journal of Archaeological Science: Reports 23: 592616.Google Scholar
Dunbar, R. I. M. 1992. Neocortex size as a constraint on group size in primates. Journal of Human Evolution 22 (6): 469–93.Google Scholar
Dunbar, R. I. M. 2003. The social brain: mind, language and society in evolutionary perspective. Annual Review of Anthropology 32: 163–81.Google Scholar
Dunsworth, H. M., Warrener, A. G., Deacon, T., Ellison, P. T. & Pontzer, H.. 2012. Metabolic hypothesis for human altriciality. Proceedings of the National Academy of Sciences USA 109 (38): 15212–16.Google Scholar
Dutton, A., Carlson, A. E., Long, A. et al. 2015. Sea-level rise due to polar ice-sheet mass loss during past warm periods. Science 349 (6244): aaa4019.Google Scholar
Duval, M., Bahain, J.-J., Falgueres, C. et al. 2015. Revisiting the ESR chronology of the Early Pleistocene hominin occupation at Vallparadís (Barcelona, Spain). Quaternary International 389: 213–23.Google Scholar
Duveau, J., Berillon, G., Verna, C., Laisné, G. & Cliquet, D.. 2019. The composition of a Neandertal social group revealed by the hominin footprints at Le Rozel (Normandy, France). Proceedings of the National Academy of Sciences USA 116 (39): 19409–14.Google Scholar
Dyble, M., Salali, G. D., Chaudhary, N. et al. 2015. Sex equality can explain the unique social structure of hunter-gatherer bands. Science 348 (6236): 796–8.Google Scholar
Early, J. D. 1985. Low forager fertility: demographic characteristic or methodological artifact? Human Biology 57 (3): 387–99.Google Scholar
Early, J. D. & Headland, T. N.. 1998. Population Dynamics of a Philippine Rain Forest People: The San Ildefonse Agta. Gainesville: University of Florida Press.Google Scholar
Eaton, J. W. & Mayer, A. J.. 1953. The social biology of very high fertility among the Hutterites: the demography of a unique population. Human Biology 25 (3): 206–64.Google Scholar
Eder, J. F. 1987. On the Road to Tribal Extinction: Depopulation, Deculturation, and Adaptive Well-Being among the Batak of the Philippines. Berkeley: University of California Press.Google Scholar
Edinborough, K. 2009. Population history and the evolution of Mesolithic arrowhead technology in South Scandinavia. In Shennan, S (ed.), Pattern and Process in Cultural Evolution. Berkley: University of California Press, pp. 191202.Google Scholar
Eixea, A. 2018. Middle Palaeolithic lithic assemblages in western Mediterranean Europe from MIS 5 to 3. Journal of Archaeological Science: Reports 21: 643–66.Google Scholar
El Zaatari, S. & Hublin, J-J. 2014. Diet of Upper Palaeolithic modern humans: evidence from microwear texture analysis. American Journal of Physical Anthropology 153: 570–81.Google Scholar
El Zaatari, S., Grine, F. E., Ungar, P. S. & Hublin, J. J.. 2011. Ecogeographic variation in Neandertal dietary habits: evidence from occlusal molar microwear texture analysis. Journal of Human Evolution 61 (4): 411–24.Google Scholar
Eller, E., Hawks, J. & Relethford, J. H.. 2004. Local extinction and recolonization, species effective population size, and modern human origins. Human Biology 81 (5/6): 805–24.Google Scholar
Ellison, P. T. 2001. On Fertile Ground: A Natural History of Human Reproduction. Cambridge, MA: Harvard University Press.Google Scholar
Ellison, P. T. 2003. Energetics and reproductive effort. American Journal of Human Biology 15: 342351.Google Scholar
Ellison, P. T. 2008. Energetics, reproductive ecology, and human evolution. PaleoAnthropology 2008: 172200.Google Scholar
Emery Thompson, M. E. 2013. Comparative reproductive energetics of human and nonhuman primates. Annual Review of Anthropology 42: 287304.Google Scholar
Ermini, L., Der Sarkissian, C., Willerslev, E. & Orlando, L.. 2015. Major transitions in human evolution revisited: a tribute to ancient DNA. Journal of Human Evolution 79: 420.Google Scholar
Estabrook, V. H. 2007. Is trauma at Krapina like all other Neandertal trauma? A statistical comparison of trauma patterns in Neandertal skeletal remains. Periodicum Biologorum 109 (4): 393400.Google Scholar
Estabrook, V. H. 2009. Sampling Bias and New Ways of Addressing the Significance of Trauma in Neandertals. Unpublished PhD thesis: University of Michigan.Google Scholar
Estabrook, V. H. 2014. Violence and warfare in the European Mesolithic and Paleolithic. In Allen, M & Jones, T. (eds.), Violence and Warfare Among Hunter Gatherers. Walnut Creek, CA: Left Coast Press, pp. 4969.Google Scholar
Estabrook, V. H. & Frayer, D. W.. 2014. Trauma in the Krapina Neandertals: violence in the Middle Palaeolithic? In Knüsel, C. & Smith, M. J. (eds.), The Routledge Handbook of the Bioarchaeology of Human Conflict. London: Routledge, pp. 6789.Google Scholar
Estalrrich, A. & Rosas, A.. 2015. Division of labor by sex and age in Neandertals: an approach through the study of activity-related dental wear. Journal of Human Evolution 80: 5163.Google Scholar
Estalrrich, A., El Zaatari, S. & Rosas, A.. 2017. Dietary reconstruction of the El Sidrón Neandertal familial group (Spain) in the context of other Neandertal and modern hunter-gatherer groups: a molar microwear texture analysis. Journal of Human Evolution 104: 1322.Google Scholar
Fabbri, F. & Mallegni, F.. 1988. Dental anthropology of the Upper Palaeolithic remains from Romito cave at Papasidero (Cosenza, Italy). Bulletins de Mémoires de la Société d’Anthropologie de Paris 5 (3): 163177.Google Scholar
Fabre., V., Condemi, S. & Degioanni, A.. 2009. Genetic evidence of geographical groups among Neanderthals. PloS One 4 (4): e5151.Google Scholar
Faivre, J. P., Discamps, E., Gravina, B., Turq, A., Guadelli, J. L. & Lenoir, M.. 2014. The contribution of lithic production systems to the interpretation of Mousterian industrial variability in south-western France: the example of Combe-Grenal (Dordogne, France). Quaternary International 350: 227–40.Google Scholar
Falcucci, A., Conard, N. J. & Peresani, M.. 2017. A critical assessment of the Protoaurignacian lithic technology at Fumane Cave and its implications for the definition of the earliest Aurignacian. PloS One 12 (12): e0189241.Google Scholar
Falguères, C., Shao, Q., Han, F. et al. 2015. New ESR and U-Series dating at Caune de l’Arago, France: a key-site for European Middle Pleistocene. Quaternary Geochronology 30: 547–53.CrossRefGoogle Scholar
Farbstein, R. 2017. Palaeolithic Central and Eastern Europe. In Insoll, T (ed.), The Oxford Handbook of Prehistoric Figurines. Oxford: Oxford University Press, pp. 681704.Google Scholar
Fareed, M. & Afzal, M.. 2017. Genetics of consanguinity and inbreeding in health and disease. Annals of Human Biology 44 (2): 99107.Google Scholar
Fay, N., De Kleine, N., Walker, B. & Caldwell, C. A.. 2019. Increasing population size can inhibit cumulative cultural evolution. Proceedings of the National Academy of Sciences USA 116 (14): 6726–31.Google Scholar
Féblot-Augustins, J. 1993. Mobility strategies in the Late Middle Palaeolithic of Central Europe and Western Europe: elements of stability and variability. Journal of Anthropological Archaeology 12: 211–65.Google Scholar
Féblot-Augustins, J. 1999. Raw material transport patterns and settlement systems in the European Lower and Middle Palaeolithic: continuity, change, and variability. In Roebroeks, W & Gamble, C (eds.), The Middle Palaeolithic Occupation of Europe. Leiden: Leiden University Press pp. 193214.Google Scholar
Féblot-Augustins, J. 2009. Revisiting European Upper Palaeolithic raw material transfers: the demise of the cultural ecological paradigm? In Adams, B. & Blades, B. S. (eds.), Lithic Materials and Palaeolithic Societies. Oxford: Wiley Blackwell, pp. 2546.Google Scholar
Fennell, K. J. & Trinkaus, E.. 1997. Bilateral femoral and tibial periostitis in the La Ferrassie 1 Neanderthal. Journal of Archaeological Science 24: 985–95.Google Scholar
Fernández López de Pablo, J., Gutiérrez-Roig, M, Gómez-Puche, M, McLaughlin, R, Silva, F & Lozano, S. 2019. Palaeodemographic modelling supports a population bottleneck during the Pleistocene-Holocene transition in Iberian. Nature Communications 10: 1872.Google Scholar
Fernández-Jalvo, Y. & Andrews, P.. 2019. Spy cave (Belgium) Neanderthals (36,000 y BP): taphonomy and peri-mortem traumas of Spy I and Spy II: Murder or accident. Quaternary Science Reviews 217: 119–29.Google Scholar
Ferring, R., Oms, O., Agustí, J. et al. 2011. Earliest human occupations at Dmanisi (Georgian Caucasus) dated to 1.85–1.78 Ma. Proceedings of the National Academy of Sciences USA 108 (26): 10432–6.Google Scholar
Feurdean, A., Perşoiu, A., Tant̹ău, T. et al. 2014. Climatic variability and associated vegetation response throughout Central and Eastern Europe (CEE) between 60 and 8 ka. Quaternary Science Reviews 106: 206–24.Google Scholar
Fewlass, H., Talamo, S., Wacker, L. et al. 2020. A 14C chronology for the Middle to Upper Palaeolithic transition at Bacho Kiro Cave, Bulgaria. Nature Ecology and Evolution 4: 794801.Google Scholar
Fiedler, L. 1990. Ein mittelpaläolithischer Hüttengrundriss aus Edertal-Buhlen in Nordhessen. Ethnographisch-Archäologische Zeitschrift 31: 6573.Google Scholar
Finlayson, C. 2004. Neanderthals and Modern Humans: An Ecological and Evolutionary Perspective. Cambridge: Cambridge University Press.Google Scholar
Finlayson, C. 2008. On the importance of coastal areas in the survival of Neanderthal populations during the late Pleistocene. Quaternary Science Reviews 27: 2246–52.Google Scholar
Finlayson, C. & Carrión, J. S.. 2007. Rapid ecological turnover and its impact on Neanderthals and other human populations. Trends in Ecology and Evolution 22 (4): 213–22.Google Scholar
Finlayson, C., Fa, D. A., Finlayson, G., Pacheco, F. G. & Rodríguez Vidal, J.. 2004. Did the moderns kill off the Neanderthals? A reply to F. d’Errico & Sánchez Goñi. Quaternary Science Reviews 23: 1205–16.Google Scholar
Finlayson, C., Giles Pacheco, F., Rodríguez-Vidal, J. et al. 2006. Late survival of Neanderthals at the southernmost extreme of Europe. Nature 443: 850–53.Google Scholar
Finlayson, C., Brown, K., Blasco, R. et al. 2012. Birds of a feather: Neanderthal exploitation of raptors and corvids. PloS One 7 (9): e45927.Google Scholar
Fiorenza, L., Benazzi, S., Tausch, J., Kullmer, O., Bromage, T. G. & Schrenk, F.. 2011. Molar macrowear reveals Neanderthal eco-geographic dietary variation. PLoS One 6 (3): e14769.Google Scholar
Fišáková, M. N. 2013. Seasonality of Gravettian sites in the Middle Danube Region and adjoining areas of Central Europe. Quaternary International 294: 120–34.Google Scholar
Fitzhugh, B., Phillips, S. C. & Gjesfjeld, E.. 2011. Modeling hunter-gatherer information networks: an archaeological case study from the Kuril Islands. In Whallon, R, Lovis, W. A & Hitchcock, R. K. (eds.), Information and Its Role in Hunter-Gatherer Bands. Los Angeles: Cotsen Institute of Archaeology Press. Ideas, Debates, and Perspectives 5, pp. 85115.Google Scholar
Fitzhugh, W. W. 1997. Biogeographical archaeology in the eastern North American Arctic. Human Ecology 25 (3): 385418.Google Scholar
Flannery, K. V. 1969. Origins and ecological effects of early domestication in Iran and the Near East. In Ucko, P. J. & Dimbleby, G. W. (eds.), The Domestication and Exploitation of Plants and Animals. Duckworth: London, pp. 73100.Google Scholar
Flas, D. 2011. The Middle to Upper Palaeolithic transition in Northern Europe: the Lincombian-Ranisian-Jerzmanowician and the issue of acculturation of the last Neanderthals. World Archaeology 43 (4): 605–27.Google Scholar
Fogarty, L & Creanza, N.. 2017. The niche construction of cultural complexity: interactions between innovations, population size and the environment. Philosophical Transactions of the Royal Society B 372: 20160428.Google Scholar
Foley, R. & Gamble, C.. 2009. The ecology of social transitions in human evolution. Philosophical Transactions of the Royal Society B: Biological Sciences 364 (1533): 3267–79.Google Scholar
Formicola, V. 1995. X-linked hypophosphatemic rickets: a probably Upper Palaeolithic case. American Journal of Physical Anthropology 98: 403–9.Google Scholar
Formicola, V. 2007. From the Sunghir children to the Romito dwarf. Current Anthropology 48 (3): 446–53.Google Scholar
Formicola, V. & Buzhilova, A. P.. 2004. Double child burial from Sunghir (Russia): pathology and inferences for Upper Palaeolithic funerary practices. American Journal of Physical Anthropology 124: 189–98.Google Scholar
Formicola, V. & Holt, B.. 2007. Resource availability and stature decrease in Upper Palaeolithic Europe. Journal of Anthropological Sciences 229: 147–55.Google Scholar
Formicola, V. & Holt, B. M.. 2015. Tall guys and fat ladies: Grimaldi’s Upper Paleolithic burials and figurines in an historical perspective. Journal of Anthropological Sciences 93: 7188.Google Scholar
Formicola, V., Frayer, D. W. & Heller, J. A.. 1990. Bilateral absence of the lesser trochanter in a Late Epigravettian skeleton from Arene Candide (Italy). American Journal of Physical Anthropology 83: 425–37.Google Scholar
Formicola, V., Pontrandolfi, A. & Svoboda, J.. 2001. The Upper Paleolithic triple burial of Dolní Věstonice: pathology and funerary behaviour. American Journal of Physical Anthropology 115: 372–9.Google Scholar
Formicola, V., Pettitt, P. B., Maggi, R. & Hedges, R.. 2005. Tempo and mode of formation of the Late Epigravettian necropolis of Arene Candide cave (Italy): direct radiocarbon evidence. Journal of Archaeological Science 32: 1598–602.Google Scholar
Fort, J., Crema, E. R. & Madella, M.. 2015. Modeling demic and cultural diffusion: an introduction. Human Biology 87 (3): 141–9.Google Scholar
Franciscus, R. G. 2009. When did the modern human pattern of childbirth arise? New insights from an old Neanderthal pelvis. Proceedings of the National Academy of Sciences USA 106 (22): 9125–6.Google Scholar
Frayer, D. W., Horton, W. A., Macchiarlli, R. & Mussi, M.. 1987. Dwarfism in an adolescent from the Italian late Upper Palaeolithic. Nature 330: 60–2.Google Scholar
Freeman, J., Byers, D. A., Robinson, E. & Kelly, R. L.. 2018. Culture process and the interpretation of radiocarbon data. Radiocarbon 60 (2): 453–67.Google Scholar
French, J. C. 2010. Double standards and ‘boundary bias’ at the European Middle-to-Upper Palaeolithic transition? An examination of the treatment of evidence for Palaeolithic habitation structures. Archaeological Review from Cambridge 25 (2): 928.Google Scholar
French, J. C. 2013. Populating the Palaeolithic: a palaeodemographic analysis of Upper Palaeolithic hunter-gatherer populations in Southwestern France. Unpublished PhD thesis, University of Cambridge.Google Scholar
French, J. C. 2015. The demography of the Upper Palaeolithic hunter-gatherers of Southwestern France: a multi-proxy approach using archaeological data. Journal of Anthropological Archaeology 39: 193209.Google Scholar
French, J. C. 2016. Demography and the Palaeolithic archaeological record. Journal of Archaeological Method and Theory 23 (1): 150–99.Google Scholar
French, J. C. 2019a. Archaeological demography as a tool for the study of women and gender in the past. Cambridge Archaeolological Journal 9: 141–57.Google Scholar
French, J. C. 2019b. The use of ethnographic data in Neanderthal archaeological research: Recent trends and their interpretative implications. Hunter Gatherer Research 4 (1): 2549.Google Scholar
French, J. C. & Chamberlain, A. T.. 2021. Demographic uniformitarianism: the theoretical basis of prehistoric demographic research and its cross-disciplinary challenges. Philosophical Transactions of the Royal Society B 376: 20190720.Google Scholar
French, J. C. & Collins, C. M.. 2015. Upper Palaeolithic population histories of Southwestern France: a comparison of the demographic signatures of 14C date distributions and archaeological site counts. Journal of Archaeological Science 55: 122–34.Google Scholar
Frisch, R. E. 1978. Population, food intake, and fertility. Science 199: 2230.Google Scholar
Frisch, R. E. 1994. The right weight: body fat, menarche and fertility. Proceedings of the Nutrition Society 53: 113–29.Google Scholar
Frisch, R. E. 2002. Female Fertility and the Body Fat Connection. Chicago: University of Chicago Press.Google Scholar
Froehle, A. W. & Churchill, S. E.. 2009. Energetic competition between Neanderthals and anatomically modern humans. Paleoanthropology 2009: 96116.Google Scholar
Froehle, A. W., Yokley, T. R. & Churchill, S. E.. 2013 . Energetics and the origin of modern humans. In Smith, F. H. & Ahern, J. C. (eds.), The Origins of Modern Humans: Biology Reconsidered (2nd ed.). Hoboken, NJ: John Wiley and Sons, pp. 285320.Google Scholar
Fu, Q., Li, H., Moorjani, P. et al. 2014. Genome sequence of a 45,000-year-old modern human from western Siberia. Nature 514: 445–50.Google Scholar
Fu, Q., Hajdinjak, M., Moldovan, O. T. et al. 2015. An early modern human from Romania with a recent Neanderthal ancestor. Nature 524: 216–19.Google Scholar
Fu, Q., Posth, C., Hajdinjak, M.. 2016. The genetic history of Ice Age Europe. Nature 534: 200–5.Google Scholar
Fuentes, O., Lucas, C. & Robert, E.. 2019. An approach to Palaeolithic networks: the question of symbolic territories and their interpretation through Magdalenian art. Quaternary International 503: 233–47.Google Scholar
Gabarda, M., Vicente, R., Martínez Valle, P. M. et al. 2016. The Lower Palaeolithic site Alto de las Picarazas (Andilla-Chelva, Valencia). Quaternary International 393 (2016): 8394.Google Scholar
Gabunia, L., Vekua, A., Lordkipanidze, D. et al. 2000. Earliest Pleistocene hominid cranial remains from Dmanisi, Republic of Georgia: taxonomy, geological setting, and age. Science 288: 1019–25.Google Scholar
Gage, T. B., DeWitte, S. & Wood, J. W.. 2012. Demography, part 1: Mortality and Migration. In Stinson, S., Bogin, B. & O’Rourke, D. (eds.), Human Biology: An Evolutionary and Biocultural Perspective (2nd ed.). Chichester: John Wiley & Sons, pp. 693755.Google Scholar
Gallivan, M. D. 2002. Measuring sedentariness and settlement population: accumulations research in the Middle Atlantic region. American Antiquity 67 (3): 535–57.Google Scholar
Gamble, C. 1982. Interaction and alliance in Palaeolithic society. Man 17 (1): 92107.Google Scholar
Gamble, C. 1986. The Palaeolithic Settlement of Europe. Cambridge: Cambridge University Press.Google Scholar
Gamble, C. 1987. Man the shoveller: alternative models for Middle Pleistocene colonisation and occupation in northern latitudes. In Soffer, O (ed.), The Pleistocene Old World: Regional Perspectives. New York,: Plenum Press, pp. 8198.Google Scholar
Gamble, C. 1996. Hominid behaviour in the Middle Pleistocene: an English perspective. In Gamble, C. S & Lawson, A. J (eds.), The English Palaeolithic Reviewed. Salisbury: Trust for Wessex Archaeology, pp. 6671.Google Scholar
Gamble, C. 1998. Palaeolithic society and the release from proximity: a network approach to intimate relations. World Archaeology 29 (3): 426–49.Google Scholar
Gamble, C. 1999. The Palaeolithic Societies of Europe. Cambridge: Cambridge University Press.Google Scholar
Gamble, C. 2009. Human display and dispersal: a case study from biotidal Britain in the Middle and Upper Pleistocene. Evolutionary Anthropology 18: 144–56.Google Scholar
Gamble, C. 2013. Settling the Earth: The Archaeology of Deep Human History. Cambridge: Cambridge University Press.Google Scholar
Gamble, C., Davies, W., Pettitt, P. & Richards, M.. 2004. Climate change and evolving human diversity in Europe during the last glacial. Philosophical Transactions of the Royal Society B 359: 243–54.Google Scholar
Gamble, C., Davies, W., Pettitt, P., Hazelwood, L. & Richards, M.. 2005. The archaeological and genetic foundations of the European population during the Late Glacial: implications for ‘agricultural thinking’. Cambridge Archaeological Journal 15 (2): 193223.Google Scholar
Garcia, J., Martínez, K. & Carbonell, E.. 2011. Continuity of the first human occupation in the Iberian Peninsula: closing the archaeological gap. Comptes Rendus Palevol 10: 279–84.Google Scholar
García-González, R., Miguel Carretero, J., Richards, M. P., Rodríguez, L. & Quam, R.. 2015. Dietary inferences through dental microwear and isotope analyses of the Lower Magdalenian individual from El Mirón Cave (Cantabria, Spain). Journal of Archaeological Science 60: 2838.Google Scholar
Gardner, J. C. & Smith, F. H.. 2006. The paleopathology of the Krapina Neandertals. Periodicum Biologorum 108 (4): 471–84.Google Scholar
Garvey, R. 2018. Current and potential roles of archaeology in the development of cultural evolutionary theory. Philosophical Transactions of the Royal Society B 373: 20170057.Google Scholar
Gaudzinski-Windheuser, S. 2011. An introduction to living structures and history of occupation at the Late Upper Palaeolithic site of Oelknitz (Thuringia, Germany). In Gaudzinski-Windheuser, S., Jöris, O., Sensburg, M., Street, M. & Turner, E. (eds.), Site-Internal Spatial Organization of Hunter-Gatherer Societies: Case Studies from the European Palaeolithic and Mesolithic. Mainz: Verlag des Römisch-Germanischen Zentralmuseums, pp. 127–39.Google Scholar
Gaudzinski-Windheuser, S. 2015. The public and private use of space in Magdalenian societies: evidence from Oelknitz 3, LOP (Thuringia, Germany). Journal of Anthropological Archaeology 40: 361–75.Google Scholar
Gaudzinski-Windheuser, S. & Jöris, O.. 2015. Contextualising the female image: symbols for common ideas and communal identity in Upper Palaeolithic societies. In Coward, F, Hosfield, R, Pope, M & Wenban-Smith, F (eds.), Settlement, Society and Cognition in Human Evolution. Landscapes in Mind. Cambridge: Cambridge University Press pp. 288314.Google Scholar
Gaudzinski-Windheuser, S. & Roebroeks, W.. 2011. On Neanderthal subsistence in last interglacial forested environments in Northern Europe. In Conard, N. J. & Richter, J. (eds.), Neanderthal Lifeways, Subsistence and Technology: One Hundred Fifty Years of Neanderthal Study. Dordrecht: Springer, pp. 6171.Google Scholar
Gautney, J. R. & Holliday, T. W.. 2015. New estimations of habitable land area and human population size at the Last Glacial Maximum. Journal of Archaeological Science 58: 103–12.Google Scholar
Ghisleni, L., Jordan, A. M. & Fioccoprile, E.. 2016. Introduction to ‘binary binds’: deconstructing sex and gender dichotomies in archaeological practice. Journal of Archaeological Method and Theory 23: 765–87.Google Scholar
Giacobini, G. 2007. Richness and diversity of burial rites in the Upper Palaeolithic. Diogenes 214: 1939.Google Scholar
Gilligan, I. 2007. Neanderthal extinction and modern human behaviour: the role of climate change and clothing. World Archaeology 39 (4): 499514.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
Gilpin, W., Feldman, M. W. & Aoki, K.. 2016. An ecocultural model predicts Neanderthal extinction through competition with modern humans. Proceedings of the National Academy of Sciences USA 113 (8): 2134–9.Google Scholar
Giot, P., Talec, L., Monnier, J.-L. & Allard, M.. 1975. Le Paléolithique supérieur du pays de Léon (Finistère). Le gisement de Beg-ar-C’hastel en Kerlouan. L’Anthropologie 79 (1): 3979.Google Scholar
Gladkih, M. I., Kornietz, N. L. & Soffer, O.. 1984. Mammoth-bone dwellings on the Russian plain. Scientific American 251 (5): 164–75.Google Scholar
Goldfield, A. E., Booton, R. & Marston, J. M.. 2019. Modeling the role of fire and cooking in the competitive exclusion of Neanderthals. Journal of Human Evolution 124: 91104.Google Scholar
Golovanova, L. V., Doronichev, V. B., Cleghorn, N. E., Koulkova, M. A., Sapelko, T. V. & Shackley, M. S.. 2010. Significance of ecological factors in the Middle to Upper Palaeolithic transition. Current Anthropology 51 (5): 655–91.Google Scholar
Goodman, M. J., Estioko-Griffin, A., Griffin, P. B. & Grove, J. S.. 1985. Menarche, pregnancy, birth spacing and menopause among the Agta women foragers of Cagayan Province, Luzon, the Philippines. Annals of Human Biology 12 (2): 169–77.Google Scholar
Goval, E., Hérrison, D., Locht, J-L & Coudenneau, A.. 2016. Levallois points and triangular flakes during the Middle Palaeolithic in northwestern Europe: considerations on the status of these pieces in the Neanderthal hunting tool kit in northern France. Quaternary International 411: 216–32.Google Scholar
Gracia, A., Arsuaga, J. L., Martínez, I. et al. 2009. Craniosynostosis in the Middle Pleistocene human cranium 14 from the Sima de los Huesos, Atapuerca, Spain. Proceedings of the National Academy of Sciences 106 (16): 6573–8.Google Scholar
Gracia-Téllez, A., Arsuaga, J. L., Martínez, I. et al. 2013. Orofacial pathology in Homo heidelbergensis: the case of Skull 5 from the Sima de los Huesos site (Atapuerca, Spain). Quaternary International 295: 8393.Google Scholar
Gravel-Miguel, C. 2016. Using Species Distribution Modeling to contextualize Lower Magdalenian social networks visible through portable art stylistic similarities in the Cantabrian region (Spain). Quaternary International 412: 112–23.Google Scholar
Gravina, B., Mellars, P. & Bronk Ramsey, C.. 2005. Radiocarbon dating of interstratified Neanderthal and early modern human occupations at the Châtelperronian type-site. Nature 438 (3): 51–6.Google Scholar
Gravina, B., Bachellerie, F., Caux, S. et al. 2018. No reliable evidence for a Neanderthal-Châtelperronian association at La Roche-à-Pierrot, Saint-Césaire. Scientific Reports 8: 15134.Google Scholar
Grayson, D. K. & Delpech, F.. 2003. Ungulates and the Middle-to-Upper Paleolithic transition at Grotte XVI (Dordogne, France). Journal of Archaeological Science 30: 1633–48.Google Scholar
Green, R. E., Krause, J., Ptak, S. E. et al. 2006. Analysis of one million base pairs of Neanderthal DNA. Nature 444 (7117): 330–6.Google Scholar
Green, R. E., Krause, J., Briggs, A. W. et al. 2010. A draft sequence of the Neanderthal genome. Science 328: 710–22.Google Scholar
Greenbaum, G., Friesem, D. E., Hovers, E., Feldman, M. W. & Kolodny, O.. 2019. Was inter-population connectivity of Neanderthals and modern humans the driver of the Upper Paleolithic transition rather than its product? Quaternary Science Reviews 217: 316–29.Google Scholar
Gretzinger, J., Molak, M., Reiter, E. et al. 2019. Large-scale mitogenomic analysis of the phylogeography of the Late Pleistocene cave bear. Scientific Reports 9 (1): 111.Google Scholar
Groucutt, H. S. 2020. Into the tangled web of culture-history and convergent evolution. In Groucutt, H. S (ed.), Culture History and Convergent Evolution: Can we Detect Populations in Prehistory? Cham, Switzerland: Springer, pp. 112.Google Scholar
Groucutt, H. S., Petraglia, M. D., Bailey, G. et al. 2015. Rethinking the dispersal of Homo sapiens out of Africa. Evolutionary Anthropology 24: 149–64.Google Scholar
Groucutt, H. W., Grün, R., Zalmout, I. S. A. et al. 2018. Homo sapiens in Arabia by 85,000 years ago. Nature Ecology and Evolution 2: 800–9.Google Scholar
Grove, M. 2009. Hunter-gatherer movement patterns: causes and constraints. Journal of Anthropological Archaeology 28: 222–33.Google Scholar
Grove, M. 2010. Logistical mobility reduces subsistence risk in hunting economies. Journal of Archaeological Science 37: 1913–21.Google Scholar
Grove, M. 2012. Scatters, patches and palimpsests: solving the contemporaneity problem. In Ruebens, K., Romanowska, I. & Bynoe, R. (eds.), Unravelling the Palaeolithic: Ten Years of Research at the Centre for the Archaeology of Human Origins (CAHO, University of Southampton). Oxford: Archeopress, pp. 153–64.Google Scholar
Grove, M. 2016. Population density, mobility, and cultural transmission. Journal of Archaeological Science 74: 7584.Google Scholar
Grove, M., Pearce, E. & Dunbar, R. I. M.. 2012. Fission-fusion and the evolution of hominin social systems. Journal of Human Evolution 62: 191200.Google Scholar
Guatelli-Steinberg, D., Spencer, C. S. & Hutchinson, D. L.. 2004. Prevalence and the duration of linear enamel hypoplasia: a comparative study of Neandertals and Inuit foragers. Journal of Human Evolution 47: 6584.Google Scholar
Guatelli-Steinberg, D., Reid, D. J., Bishop, T. A. & Larsen, C. S.. 2005. Anterior tooth growth periods in Neandertals were comparable to those of modern humans. Proceedings of the National Academy of Sciences USA 102 (40): 14197–202.Google Scholar
Guatelli-Steinberg, D., Buzhilova, A. P. & Trinkaus, E.. 2013. Developmental stress and survival among the Mid Upper Palaeolithic Sunghir children: dental enamel hypoplasias of Sunghir 2 and 3. International Journal of Osteoarchaeology 23: 421–31.Google Scholar
Guatelli-Steinberg, D., Stinespring-Harris, A., Reid, D. J., Larsen, C. S., Hutchinson, D. L. & Smith, T. M.. 2014. Chronology of linear enamel hypoplasia formation in the Krapina Neanderthals. PaleoAnthropology 2014: 431–45.Google Scholar
Günther, T. & Jakobsson, M. 2016. Genes mirror migrations and cultures in prehistoric Europe—a population genomic perspective. Current Opinion in Genetics and Development 41: 115–23.Google Scholar
Gurven, M. D & Davison, R. J.. 2019. Periodic catastrophes over human evolutionary history are necessary to explain the forager population paradox. Proceedings of the National Academy of Sciences USA 116: 19.Google Scholar
Gurven, M. & Kaplan, H.. 2007. Longevity among hunter-gatherers: a cross-cultural examination. Population and Development Review 33 (2): 321–65.Google Scholar
Guy, H., Masset, C. & Baud, C-A. 1997. Infant taphonomy. International Journal of Osteoarchaeology 7: 221–29.Google Scholar
Haesaerts, P., Damblon, F., Neugebauer-Maresch, C. & Einwögerer, T.. 2016. Radiocarbon chronology of the Late Palaeolithic loess site of Kammern-Grubgraben (Lower Austria). Archaeologia Austriaca 100: 271–7.Google Scholar
Hajdinjak, M., Fu, Q., Hübner, A. et al. 2018. Reconstructing the genetic history of late Neanderthals. Nature 555: 652–6.Google Scholar
Ham, E., Underdown, S. J. & Houldcroft, C. J.. 2019. The relative roles of maternal survival and inter-personal violence as selection pressures on the persistence of Neanderthal hypercoagulability alleles in modern Europeans. Annals of Human Biology 46 (2): 99108.Google Scholar
Hames, R. & Draper, P.. 2004. Women’s work, child care and helpers-at-the-nest in a hunter-gatherer society. Human Nature 15 (4): 319–41.Google Scholar
Hamilton, M. J & Walker, R. S.. 2018. A stochastic density-dependent model of long-term population dynamics in hunter-gatherer populations. Evolutionary Ecology Research 19: 85102.Google Scholar
Hamilton, M. J., Milne, B. T., Walker, R. S., Burger, O. & Brown, J. H.. 2007. The complex structure of hunter-gatherer social networks. Proceedings of the Royal Society B 274: 2195–202.Google Scholar
Hammer, M. L. A. & Foley, R. A.. 1996. Longevity and life history in hominid evolution. Human Evolution 11: 61–6.Google Scholar
Hardy, B. L., Moncel, M. H., Kerfant, C., Lebon, M., Bellot-Gurlet, L. & Mélard, N.. 2020. Direct evidence of Neanderthal fibre technology and its cognitive and behavioral implications. Scientific Reports 10 (1): 19.Google Scholar
Hardy, K., Buckley, S. & Huffman, M.. 2013. Neanderthal self-medication in context. Antiquity 87 (337): 873–8.Google Scholar
Harpending, H. 1994. Infertility and forager demography. American Journal of Physical Anthropology 93: 385–90.Google Scholar
Harpending, H. & Wandsnider, L.. 1982. Population structures of the Ghanzi and Ngamiland !Kung. In Crawford, M. H & Wandsnider, L. (eds.), Current Developments in Anthropological Genetics, Vol. 2. New York: Plenum Press, pp. 2950.Google Scholar
Harris, E. E. 2017. Demic and cultural diffusion in prehistoric Europe in the age of ancient genomes. Evolutionary Anthropology 26 (5): 228–41.Google Scholar
Harris, K. & Nielsen, R.. 2016. The genetic cost of Neanderthal introgression. Genetics 203 (2): 881–91.Google Scholar
Harvati, K., 2003. The Neanderthal taxonomic position: models of intra-and inter-specific craniofacial variation. Journal of Human Evolution 44 (1): 107–32.Google Scholar
Harvati, K., Röding, C., Bosman, A. M. et al. 2019. Apidima Cave fossils provide earliest evidence of Homo sapiens in Eurasia. Nature 571 (7766): 500–4.Google Scholar
Hassan, F. 1975. Determination of the size, density, and growth rate of hunting-gathering populations. In Polgar, S. (ed.), Population, Ecology, and Social Evolution. The Hague: Mouton, pp. 2752.Google Scholar
Hassan, F. 1981. Demographic Archaeology. New York: Academic Press.Google Scholar
Hauck, T. C., Rethmeyer, J., Rentzel, P. et al. 2016. Neanderthals or early modern humans? A revised 14C chronology and geoarchaeological study of the Szeletian sequence in Szeleta cave (Kom. Borsod-Abaúj-Zemplén) in Hungary. Archäologisches Korrespondenzblatt 46: 271–90.Google Scholar
Haws, J. A., Benedetti, M. M., Talamo, S. et al. 2020. The early Aurignacian dispersal of modern humans into westernmost Eurasia. Proceedings of the National Academy of Sciences USA 117 (41): 25414–22.Google Scholar
Hawkes, K. & O’Connell, J. F.. 2005. How old is human longevity? Journal of Human Evolution 49: 650–3.Google Scholar
Hawkes, K., O’Connell, J. & Blurton Jones, N.. 2018. Hunter-gatherer studies and human evolution: a very selective review. American Journal of Physical Anthropology 165: 777800.Google Scholar
Hawks, J. 2008. From genes to numbers: effective population sizes in human evolution. In Bocquet-Appel, J.-P (ed.), Recent Advances in Palaeodemography: Data, Techniques, Patterns. Dordrecht: Springer, pp. 930.Google Scholar
Hawks, J. 2012. Dynamics of genetic and morphological variability within Neandertals. Journal of Anthropological Sciences 90: 8197.Google Scholar
Hawks, J. 2017. Neanderthals and Denisovans as biological invaders. Proceedings of the National Academy of Sciences USA 114 (37): 9761–3.Google Scholar
Hayden, B. 1972. Population control among hunter/gatherers. World Archaeology 4 (2): 205–21.Google Scholar
Hayden, B. 2012. Neanderthal social structure? Oxford Journal of Archaeology 31 (1): 126.Google Scholar
Head, M. J. & Gibbard, P. L.. 2005. Early-Middle Pleistocene transitions: an overview and recommendation for the defining boundary. Geological Society, London, Special Publications 247: 118.Google Scholar
Headland, T. N. 1989. Population decline in a Philippine Negrito hunter-gatherer society. American Journal of Human Biology 1: 5972.Google Scholar
Heled, J. & Drummond, A. J.. 2008. Bayesian inference of population size history from multiple loci. BMC Evolutionary Biology 8 (1): 289.Google Scholar
Hemming, S. R. 2004. Heinrich events: massive late Pleistocene detritus layers of the North Atlantic and their global climatic imprint. Reviews of Geophysics 42: 143.Google Scholar
Henrich, J. 2004. Demography and cultural evolution: how adaptive cultural processes can produce maladaptive losses: the Tasmanian case. American Antiquity 69 (2): 197214.Google Scholar
Henry, D. 2012. The palimpsest problem, hearth pattern analysis, and Middle Palaeolithic site structure. Quaternary International 247: 246–66.Google Scholar
Henry, D. O., Hietrala, H. J., Rosen, A. M., Demidenko, Y. E., Usik, V. I. & Armagan, T. L.. 2004. Human behavioural organization in the Middle Paleolithic: were Neanderthals different? American Anthropologist 106 (1): 1731.Google Scholar
Henry-Gambier, D. 2002. Les fossiles de Cro-Magnon (Les Eyzies-de-Tayac, Dordogne). Nouvelles données sur leur position chronologique et leur attribution culturelle. Bulletins et Mémoires de la Société d’Anthropologie de Paris 14: 89112.Google Scholar
Henry-Gambier, D. 2008. Comportement des populations d’Europe au Gravettien: pratiques funéraires et interprétations. Paléo 20: 399438.Google Scholar
Henshilwood, C. S. & Marean, C. W.. 2003. The origin of modern human behaviour: critique of the models and their test implications. Current Anthropology 44: 625–52.Google Scholar
Hérisson, D., Airvaux, J., Lenoble, A., Richter, D., Claud, E. & Primault, J.. 2016a. Between the northern and southern regions of western Europe: the Acheulean site of La Grande Vallée (Colombiers, Vienne, France). Quaternary International 411: 108–31.Google Scholar
Hérisson, D., Brenet, M., Cliquet, D. et al. 2016b. The emergence of the Middle Palaeolithic in north-western Europe and its southern fringes. Quaternary International 411: 233–83.Google Scholar
Hershkovitz, I., Weber, G. W., Quam, R.. 2018. The earliest modern humans outside Africa. Science 359: 456–9.Google Scholar
Herzlinger, G. & Goren-Inbar, N.. 2019. Do a few tools necessarily mean a few people? A techno-morphological approach to the question of group size at Gesher Benot Ya’aqov, Israel. Journal of Human Evolution 128: 4558.Google Scholar
Hewlett, B. L. & Hewlett, B. S.. 2012. Hunter-gatherer adolescence. In Hewlett, B. L. (ed.), Adolescent Identity: Evolutionary, Cultural, and Developmental Perspectives. New York: Routledge, pp. 73101.Google Scholar
Hewlett, B. S. 1991a. Demography and childcare in preindustrial societies. Journal of Anthropological Research 47 (1): 137.Google Scholar
Hewlett, B. S. 1991b. Intimate Fathers. The Nature and Context of Aka Pygmy Paternal Infant Care. Ann Arbor: University of Michigan Press.Google Scholar
Higham, T., Jacobi, R., Julien, M. et al. 2010. Chronology of the Grotte du Renne (France) and implications for the context of ornaments and human remains within the Châtelperronian. Proceedings of the National Academy of Sciences USA 107 (47): 20234–9.Google Scholar
Higham, T., Compton, T., Stringer, C. et al. 2011. The earliest evidence for anatomically modern humans in northwestern Europe. Nature 479 (7374): 521–4.Google Scholar
Higham, T., Basell, L., Jacobi, R., Wood, R., Bronk Ramsey, C. & Conard, N. J.. 2012. Testing models for the beginnings of the Aurignacian and the advent of figurative art and music: The radiocarbon chronology of Geißenklösterle. Journal of Human Evolution 62 (6): 664–76.Google Scholar
Higham, T., Douka, K., Wood, R. et al. 2014. The timing and spatiotemporal patterning of Neanderthal disappearance. Nature 512: 306–9.Google Scholar
Hill, K. & Hurtado, A. M.. 1996. Ache Life History: The Ecology and Demography of a Foraging People. New York: Aldine de Gruyter.Google Scholar
Hill, K., Hurtado, A. M. & Walker, R. S.. 2007. High adult mortality among Hiwi hunter-gatherers: implications for human evolution. Journal of Human Evolution 52: 443–54.Google Scholar
Hill, K. R., Walker, R. S., Božičević, M. et al. 2011. Co-residence patterns in hunter-gatherer societies show unique human social structure. Science 331: 1286–9.Google Scholar
Hillson, S. W., Franciscus, R. G., Holliday, T. W. & Trinkaus, E.. 2006. The ages at death. In Trinkaus, E. & Svoboda, J. (eds.), Early Modern Human Evolution in Central Europe. The People of Dolní Vĕstonice and Pavlov. Oxford: Oxford University Press, pp. 3145.Google Scholar
Hockett, B. & Haws, J. A.. 2003. Nutritional ecology and diachronic trends in Paleolithic diet and health. Evolutionary Anthropology 12: 211–16.Google Scholar
Hockett, B. & Haws, J. A.. 2005. Nutritional ecology and the human demography of Neandertal extinction. Quaternary International 137: 2134.Google Scholar
Hoffecker, J. F. 2002a. Desolate Landscapes: Ice-age Settlement in Eastern Europe. New Brunswick, NJ: Rutgers University Press.Google Scholar
Hoffecker, J. F. 2002b. The Eastern Gravettian ‘Kostenki culture’ as an Arctic adaptation. Anthropological Papers of the University of Alaska 2 (1): 115–36.Google Scholar
Hoffecker, J. F. 2009. The spread of modern humans in Europe. Proceedings of the National Academy of Sciences USA 106 (38): 16040–5.Google Scholar
Hoffmann, D. L, Standish, C. D., García-Diez, M. et al. 2018. U-Th dating of carbonate crusts reveals Neandertal origin of Iberian cave art. Science 359 (6378): 912–15.Google Scholar
Hohenstein, U. T., Di Nucci, A. & Moigne, A. M.. 2009. Mode de vie à Isernia La Pineta (Molise, Italie). Stratégie d’exploitation du Bison schoetensacki par les groupes humains au Paléolithique inférieur. L’Anthropologie 113 (1): 96110.Google Scholar
Holliday, T. W. 1997. Postcranial evidence of cold adaptation in European Neandertals. American Journal of Physical Anthropology 104: 245–58.Google Scholar
Holliday, T. W., Gautney, J. R. & Friedl, L.. 2014. Right for the wrong reasons. Reflections on modern human origins in the post-Neanderthal genome era. Current Anthropology 55 (6): 696724.Google Scholar
Holt, B. 2003. Mobility in Upper Paleolithic and Mesolithic Europe: evidence from the lower limb. American Journal of Physical Anthropology 122: 200–15.Google Scholar
Holt, B. M. & Formicola, V.. 2008. Hunters of the Ice Age: the biology of Upper Paleolithic people. Yearbook of Physical Anthropology 51: 7099.Google Scholar
Holt-Lunstad, J., Smith, T. B. & Layton, J. B.. 2010. Social relationships and mortality risk: a meta-analytic review. PLoS Medicine 7 (7): e1000316.Google Scholar
Hopkinson, T. 2007. The transition from the Lower to the Middle Palaeolithic in Europe and the incorporation of difference. Antiquity 81: 294307.Google Scholar
Hopkinson, T., Nowell, A. & White, M.. 2013. Life histories, metapopulation ecology, and innovation in the Acheulian. PaleoAnthropology 2013: 6176.Google Scholar
Hoppa, R. D. & Vaupel, J. W.. 2002. The Rostock Manifesto for paleodemography: the way from stage to age. In Hoppa, R. D. & Vaupel, J. W. (eds.), Paleodemography. Age Distributions from Skeletal Samples. Cambridge: Cambridge University Press, pp. 928.Google Scholar
Horan, R. D., Bulte, E. & Shogren, J. F.. 2005. How trade saved humanity from biological exclusion: an economic theory of Neanderthal extinction. Journal of Economic Behavior & Organization 58 (1): 129.Google Scholar
Hortolà, P. & Martínez-Navarro, B.. 2013. The Quaternary megafaunal extinction and the fate of Neanderthals: an integrative working hypothesis. Quaternary International 295: 6972.Google Scholar
Hosfield, R. 2011. The British Lower Palaeolithic of the early Middle Pleistocene. Quaternary Science Reviews 30: 1486–510.Google Scholar
Hosfield, R. 2016. Walking in a winter wonderland? Strategies for Early and Middle Pleistocene survival in midlatitude Europe. Current Anthropology 57 (5): 653–82.Google Scholar
Hosfield, R. 2020. The Earliest Europeans. A Year in the Life: Seasonal Survival Strategies in the Lower Palaeolithic. Oxford: Oxbow.Google Scholar
Hosfield, R. & Cole, J.. 2018. Early hominins in north-west Europe: a punctuated long chronology? Quaternary Science Reviews 190: 148–60.Google Scholar
Hosfield, R., Cole, J. & McNabb, J.. 2018. Less of a bird’s song than a hard rock ensemble. Evolutionary Anthropology 27: 920.Google Scholar
Houldcroft, C. J. & Underdown, S. J.. 2016. Neanderthal genomics suggests a Pleistocene time frame for the first epidemiologic transition. American Journal of Physical Anthropology 160 (3): 379–88.Google Scholar
Housley, R. A., Gamble, C. S., Street, M. & Pettitt, P.. 1997. Radiocarbon evidence for the Lateglacial human recolonisation of northern Europe. Proceedings of the Prehistoric Society 63 (25): 2564.Google Scholar
Howell, F. C. 1996. Thoughts on the study and interpretation of the human fossil record. In Meikle, W. E., Howell, F. C. & Jablonski, N. G. (eds.), Contemporary Issues in Human Evolution. California Academy of Sciences Memoir 21. San Francisco, pp. 145.Google Scholar
Howell, N. 1976. Toward a uniformitarian theory of human paleodemography. Journal of Human Evolution 5: 2540Google Scholar
Howell, N. 1979. Demography of the Dobe !Kung. New York: Academic Press.Google Scholar
Howell, N. 1982. Village composition implied by a paleodemographic life table: the Libben site. American Journal of Physical Anthropology 59: 263–69.Google Scholar
Howell, N. 2010. Life histories of the Dobe !Kung: Food, Fatness, and Well-Being Over the Life Span. Berkeley: University of California Press.Google Scholar
Hublin, J.-J. 2015. The modern human colonization of western Eurasia: when and where? Quaternary Science Reviews 118: 194210.Google Scholar
Hublin, J.-J. & Roebroeks, W.. 2009. Ebb and flow or regional extinctions? On the character of Neandertal occupation of northern environments? Comptes Rendus Palvol 8: 503–9.Google Scholar
Hublin, J.-J., Talamo, S., Julien, M. et al. 2012. Radiocarbon dates from the Grotte du Renne and Saint Césaire support a Neanderthal origin for the Châtelperronian. Proceedings of the National Academy of Sciences USA 109 (46): 18743–8.Google Scholar
Hublin, J.-J., Neubauer, S. & Gunz, P.. 2015. Brain ontogeny and life history in Pleistocene hominins. Philosophical Transactions of the Royal Society B: 370: 20140062.Google Scholar
Hublin, J.-J., Ben-Ncer, A., Bailey, S. E. et al. 2017. New fossils from Jebel Irhoud (Morocco) and the pan-African origin of Homo sapiens. Nature 546: 289–92.Google Scholar
Hublin, J.-J., Sirakov, N., Aldeias, V. et al. 2020. Initial Upper Palaeolithic Homo sapiens from Bacho Kiro Cave, Bulgaria. Nature 581: 299302.Google Scholar
Huff, C. D., Xing, J., Rogers, A. R., Witherspoon, D. & Jorde, L. B.. 2010. Mobile elements reveal small population size in the ancient ancestors of Homo sapiens. Proceedings of the National Academy of Sciences USA 107 (5): 2147–52.Google Scholar
Hughes, P. D. & Gibbard, P. L.. 2015. A stratigraphical basis for the Last Glacial Maximum (LGM). Quaternary International 383: 174–85.Google Scholar
Huguet, R., Saladié, P., Cáceres, I. et al. 2013. Successful subsistence strategies of the first humans in south-western Europe. Quaternary International 295: 168–82.Google Scholar
Hull, K. L. 2011. Thinking small: hunter-gatherer demography and culture change. In Sassaman, K. E & Holly, D. H. Jr. (eds.), Hunter-Gatherer Archaeology as Historical Process. Tuscon: The University of Arizona Press, pp. 3454.Google Scholar
Hurtado, A. M. & Hill, K. R.. 1987. Early dry season subsistence ecology of Cuiva (Hiwi) foragers of Venezuela. Human Ecology 15 (2): 163–87.Google Scholar
Hussain, S. T. & Floss, H.. 2016. Streams as entanglement of nature and culture: European Upper Paleolithic river systems and their role as features of spatial organization. Journal of Archaeological Method and Theory 23 (4): 1162–218.Google Scholar
Hutchinson, D. L., Larsen, C. S. & Choi, I.. 1997. Stressed to the max? Physiological perturbation in the Krapina Neandertals. Current Anthropology 38 (5): 904–14.Google Scholar
Iakovleva, L. 2015. The architecture of mammoth bone circular dwellings of the Upper Palaeolithic settlements in Central and Eastern Europe and their socio-symbolic meanings. Quaternary International 359–60: 324–34.Google Scholar
Iakovleva, L. 2016. Mezinian landscape system (Late Upper Palaeolithic of Eastern Europe). Quaternary International 412: 415.Google Scholar
Iakovleva, L. & Djindjian, F.. 2005. New data on mammoth bone settlements of Eastern Europe in the light of the new excavations of the Gontsy site (Ukraine). Quaternary International 126–128: 195207.Google Scholar
Iakovleva, L. & Djindjian, F.. 2018. The mammoth bone dwellings of the Upper Palaeolithic settlement of Gintsi (Ukraine: A first synthesis). Археологія 4 (2018): 8694.Google Scholar
Iakovleva, L. F, Djindjian, E. N. Maschenko, S. Konik, A-M. Moigne, . 2012. The later Upper Palaeolithic site of Gontsy (Ukraine): a reference for the reconstruction of the hunter-gatherer system based on a mammoth economy. Quaternary International 255: 8693.Google Scholar
Isaac, G. L. L. 1972. Early phases of human behaviour: models in Lower Palaeolithic archaeology. In Clarke, D. L. (ed.), Models in Archaeology. London: Methuen, pp. 167–99.Google Scholar
Isler, K. & van Schaik, C. P.. 2012. How our ancestors broke through the gray ceiling: comparative evidence for cooperative breeding in early Homo. Current Anthropology 53 (Supplement 6): S453–65.Google Scholar
Ivanova, S. 2016. A route through the Balkans and implications for the earliest settlement of Europe. In Harvati, K. & Roksandic, M. (eds.), Paleoanthropology of the Balkans and Anatolia. Dordrecht: Springer, pp. 187211.Google Scholar
Ivanovaitė, L. & Riede, F.. 2018. The Final Palaeolithic hunter-gatherer colonisation of Lithuania in light of recent palaeoenvironmental research. Open Quaternary 4 (4): 121.Google Scholar
Jacobi, R. M. 2007. A collection of Early Upper Palaeolithic artefacts from Beedings, near Pulborough, West Sussex, and the context of similar finds from the British Isles. Proceedings of the Prehistoric Society 73: 229325.Google Scholar
Jacobi, R. M. & Higham, T. F. G. 2011. The British Earlier Upper Palaeolithic: settlement and chronology. In Ashton, N. M., Lewis, S. G. & Stringer, C. B. (eds.), The Ancient Human Occupation of Britain. Developments in Quaternary Science 14, pp. 181222. www.sciencedirect.com/bookseries/developments-in-quaternary-sciences/vol/14/suppl/CGoogle Scholar
Jacobs, Z., Li, B., Shunkov, M. V. et al. 2019. Timing of archaic hominin occupation of Denisova Cave in southern Siberia. Nature 565: 594–9.Google Scholar
Jaouen, K., Richards, M. P., Le Cabac, A. et al. 2019. Exceptionally high δ15N values in collagen single amino acids confirm Neandertals as high-trophic level carnivores. Proceedings of the National Academy of Sciences USA 116 (11): 4928–33.Google Scholar
Jasienska, G. 2003. Energy metabolism and the evolution of reproductive suppression in the human female. Acta Biotheoretica 51: 118.Google Scholar
Jaubert, J., Verheyden, S., Genty, D. et al. 2016. Early Neanderthal constructions deep in Bruniquel cave in southwestern France. Nature 534: 111–14.Google Scholar
Jennings, R., Finlayson, C., Fa, D. & Finlayson, G.. 2011. Southern Iberia as a refuge for the last Neanderthal populations. Journal of Biogeography 38: 18731885.Google Scholar
Jiménez-Arenas, J. M., Manuel Santonja, M. B. & Palmqvist, P.. 2011. The oldest handaxes in Europe: fact or artefact? Journal of Archaeological Science 38 (12): 3340–9.Google Scholar
Jobling, M. A. 2012. The impact of recent events on human genetic diversity. Philosophical Transactions of the Royal Society B 367: 793–9.Google Scholar
Jochim, M. 1983. Palaeolithic cave art in ecological perspective. In Bailey, G (ed.), Hunter-Gatherer Economy in Prehistory. Cambridge: Cambridge University Press, pp. 212–19.Google Scholar
Jochim, M. 1987. Late Pleistocene refugia in Europe. In Soffer, O. (ed.), The Pleistocene Old World: Regional Perspectives. New York: Plenum Press, pp. 317–31.Google Scholar
Johannsen, N. N., Larson, G., Meltzer, D. J. & Vander Linden, M.. 2017. A composite window into human history: better integration of ancient DNA studies with archaeology promises deeper insights. Science 356: 1118–20.Google Scholar
Johnson, G. A. 1978. Information sources and the development of decision-making organizations. In Redman, C. L (ed.), Social Archaeology: Beyond Subsistence and Dating: New York: Academic Press, pp. 87112.Google Scholar
Jost, A., Lunt, D., Kageyama, M. et al. 2005. High resolution simulations of the last glacial maximum climate over Europe: a solution to discrepancies with continental palaeoclimatic reconstructions? Climate Dynamics 24: 577–90.Google Scholar
Juric, I., Aeschbacher, S & Coop, G. 2016. The strength of selection against Neanderthal introgression. PLOS Genetics 12 (11): e1006340.Google Scholar
Kahlke, R-D., García, N., Kostopoulos, D. S. et al. 2011. Western Palaeartic palaeoenvironmental conditions during the Early and early Middle Pleistocene inferred from large mammal communities, and implications for hominin dispersal in Europe. Quaternary Science Reviews 30: 1368–95.Google Scholar
Kaminská, L. 2016. Gravettian and Epigravettian lithics in Slovakia. Quaternary International 406: 144–65.Google Scholar
Kaminská, L., Kozłowski, J. K. & Škrdla, P.. 2011. New approach to the Szeletian-chronology and cultural variability. Eurasian Prehistory 8 (1–2): 2949.Google Scholar
Kaplan, J. O., Pfeiffer, M., Kolen, J. C. A. & Davis, B. A. S.. 2016. Large scale anthropogenic reduction of forest cover in last glacial maximum Europe. PloS One 11 (11): e0166726.Google Scholar
Karavanić, I., Vukosavljević, N., Janković, I., Ahern, J. C. & Smith, F. H.. 2018. Paleolithic hominins and settlement in Croatia from MIS 6 to MIS 3: research history and current interpretations. Quaternary International 494: 152–66.Google Scholar
Karlin, C. & Julien, M.. 2019. An autumn at Pincevent (Seine-et-Marne, France): refitting for an ethnographic approach of a Magdalenian settlement. Archaeological and Anthropological Sciences 11: 4437–65.Google Scholar
Keckler, C. N. W. 1997. Catastrophic mortality in simulations of forager age-at-death: where did all the humans go? In Paine, R. R (eds.), Integrating Archaeological Demography: Multidisciplinary Approaches to Prehistoric Population. Centre for Archaeological Investigations, Southern Illinois University, Carbondale, pp. 205–28.Google Scholar
Kehl, M., Burow, C., Hilgers, A. et al. 2013. Late Neanderthals at Jarama VI (central Iberia)? Quaternary Research 80: 218–34.Google Scholar
Kelly, R. L. 2003. Colonization of new land by hunter-gatherers. In Rockman, M. & Steele, J. (eds.), Colonization of Unfamiliar Landscapes: the Archaeology of Adaptation. London: Routledge, pp. 4458.Google Scholar
Kelly, R. L. 2013a. The Lifeways of Hunter-Gatherers: The Foraging Spectrum. Cambridge: Cambridge University Press.Google Scholar
Kelly, R. L. 2013b. From the peaceful to the warlike: ethnographic and archaeological insights into hunter-gatherer warfare and homicide. In Fry, D. P. (ed.), War, Peace, and Human Nature: The Convergence of Evolutionary and Cultural Views: Oxford: Oxford University Press, pp. 151–67.Google Scholar
Kempe, M. & Mesoudi, A.. 2014. An experimental demonstration of the effect of group size on cultural accumulation. Evolution and Human Behavior 25: 285–90.Google Scholar
Kennedy, G. E. 2003. Palaeolithic grandmothers? Life history and early Homo. Journal of the Royal Anthropological Institute 9: 549–72.Google Scholar
Key, F. M., Posth, C., Krause, J., Herbig, A. & Bos, K. I.. 2017. Mining metagenomics data sets for ancient DNA: recommended protocols for authentication. Trends in Genetics 33 (8): 508–20.Google Scholar
Kimball, L. R., Coffey, T. S., Faulks, N. R., Dellinger, S. E., Karas, N. M. & Hidjrati, N.. 2017. A multi-instrument study of microwear polishes on Mousterian tools from Weasel Cave (Myshtulagty Lagat), Russia. Lithic Technology 42 (2–3): 6176.Google Scholar
Kissel, M. & Kim, N. C.. 2018. Emergent Warfare in our Evolutionary Past. London: Routledge.Google Scholar
Kitagawa, K., Julien, M-A, Krotova, O. et al. 2018. Glacial and post-glacial adaptations of hunter-gatherers: investigating the late Upper Paleolithic and Mesolithic subsistence strategies in the southern steppe of Eastern Europe. Quaternary International 465: 192209.Google Scholar
Klaric, L. 2007. Regional groups in the European Middle Gravettian: a reconsideration of the Rayssian technology. Antiquity 81 (311): 176–90.Google Scholar
Klein, R. G. 1973. Ice-Age Hunters of the Ukraine. Chicago: University of Chicago Press.Google Scholar
Klein, R. G. 2000. Archeology and the evolution of human behavior. Evolutionary Anthropology 9 (1): 1736.Google Scholar
Klein, R. G. 2017. Language and human evolution. Journal of Neurolinguistics 43: 204–21.Google Scholar
Klein, R. G. & Steele, T. E.. 2013. Archaeological shellfish size and later human evolution in Africa. Proceedings of the National Academy of Sciences USA 110 (27): 10910–15.Google Scholar
Kline, M. A. & Boyd, R.. 2010. Population size predicts technological complexity in Oceania. Proceedings of the Royal Society B 277: 2559–64.Google Scholar
Koenigswald, W. V. & Kolfschoten, T. van. 1996. The Mimomys-Arvicola boundary and the enamel thickness quotient (SDQ) of Arvicola as stratigraphic markers in the Middle Pleistocene. In Turner, C (ed.), The Early Middle Pleistocene in Europe. Rotterdam: Balekema, pp. 211–26.Google Scholar
Kolen, J. 1999. Hominids without homes: On the nature of Middle Palaeolithic settlement in Europe. In Roebroeks, W & Gamble, C. (eds.), The Middle Palaeolithic Occupation of Europe. Leiden: University of Leiden Press, pp. 139–75.Google Scholar
Kolobova, K. A., Roberts, R. G., Chabai, V. P. et al. 2020. Archaeological evidence for two separate dispersals of Neanderthals into southern Siberia. Proceedings of the National Academy of Sciences USA 117 (6): 2879–85.Google Scholar
Kolodny, O. & Feldman, M. W.. 2017. A parsimonious neutral model suggests Neanderthal replacement was determined by migration and random species drift. Nature Communications 8 (1): 113.Google Scholar
Kovacevic, M., Shennan, S., Vanhaeren, M., d’Errico, F & Thomas, M. G. 2015. Simulating geographical variation in material culture: were early modern humans in Europe ethnically structured? In Mesoudi, A & Aoki, K (eds.), Learning Strategies and Cultural Evolution during the Palaeolithic. Dordrecht: Springer, pp. 103–20.Google Scholar
Kozłowski, J. K. 2014. Middle Palaeolithic variability in central Europe: Mousterian vs Micoquian. Quaternary International 326: 344–63.Google Scholar
Kozłowski, S. K., Połtowicz-Bobak, M., Bobak, D. & Terberger, T.. 2012. New information from Maszycka Cave and the Late Glacial recolonisation of Central Europe. Quaternary International 272: 288–96.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 (4): 472–97.Google Scholar
Kramer, K. L & Greaves, R. D.. 2011. Postmarital residence and bilateral kin associations among hunter-gatherers. Human Nature 22: 4163.Google Scholar
Kramer, K. L. & Greaves, R. D.. 2007. Changing patterns of infant mortality and maternal fertility among Pumé foragers and horticulturalists. American Anthropologist 109 (4): 713–26.Google Scholar
Kramer, K. L., Schacht, R. & Bell, A.. 2017. Adult sex ratios and partner scarcity among hunter-gatherers: implications for dispersal patterns and the evolution of human sociality. Philosophical Transactions of the Royal Society B 372: 20160316.Google Scholar
Kranioti, E. F., Grigorescu, D. & Harvati, K.. 2019. State of the art forensic techniques reveal evidence of interpersonal violence ca. 30,000 years ago. PloS One 14 (7): e0216718.Google Scholar
Krause, J. & Pääbo, S.. 2016. Genetic time travel. Genetics 203 (1): 912.Google Scholar
Krause, J., Fu, Q., Good, J. M. et al. 2010a. The complete mitochrondrial DNA genome of an unknown hominin from southern Siberia. Nature 464: 894–97.Google Scholar
Krause, J., Briggs, A. W., Kircher, M. et al. 2010b. A complete mtDNA genome of an early modern human from Kostenki, Russia. Current Biology 20: 231–6.Google Scholar
Kretschmer, I. 2015. Demographische Untersuchungen zu Bevölkerungsdichten, Mobilität und Landnutzung im späten Jungpaläolithikum. Kölner Studien zur Prähistorischen Archäologie 6. Rahden/Westf.: Verlag Marie Leidorf.Google Scholar
Kretschmer, I. 2019. Demographic studies of hunters and gatherers in the European Late Upper Palaeolithic. In Eriksen, B. V, Harris, S, Rensink, E. (eds.), The Final Palaeolithic of Northern Eurasia. Proceedings of the Amersfoort, Schleswig and Burgos UISPP Commission Meetings Schriften des Museums für Archäologie Schloss Gottorf – Ergänzungsreihe Band 13, pp. 231245.Google Scholar
Kretzoi, M. & Dobosi, V.. 1990. Vértesszölös: Site, Man and Culture. Budapest: Akadémiai Kiadó.Google Scholar
Kuhlwilm, M., Gronau, I., Hubisz, M. J. et al. 2016. Ancient gene flow from early modern humans into Eastern Neanderthals. Nature 530: 429–35.Google Scholar
Kuhn, S. L. 2012. Emergent patterns of creativity and innovation in early technologies. Developments in Quaternary Science 16: 6987.Google Scholar
Kuhn, S. L. 2013a. Roots of the Middle Palaeolithic in Eurasia. Current Anthropology 54 (8): S255S268.Google Scholar
Kuhn, S. L. 2013b. Cultural transmission, institutional continuity and the persistence of the Mousterian. In Akazawa, T, Yoshihiro, N & Aoki, K. (eds.), Dynamics of Learning in Neanderthals and Modern Humans, Volume 1: Cultural Perspectives. Dordrecht: Springer, pp. 105–13.Google Scholar
Kuhn, S. L. 2019. Initial Upper Paleolithic: A (near) global problem and a global opportunity. Archaeological Research in Asia 17: 28.Google Scholar
Kuhn, S. L. & Stiner, M. C.. 2001. The antiquity of hunter-gatherers. In Panter-Brick, C., Layton, R. H. & Rowley-Conwy, P. (eds.), Hunter-Gatherers: An Interdisciplinary Perspective. Cambridge: Cambridge University Press, pp. 99129.Google Scholar
Kuhn, S. L. & Stiner, M. C.. 2006. What’s a mother to do? The division of labor among Neandertals and Modern humans in Eurasia. Current Anthropology 47 (6): 953–80.Google Scholar
Kuhn, S. L. & Stiner, M. C.. 2007. Paleolithic ornaments: implications for cognition, demography and identity. Diogenes 54 (2): 40–8.Google Scholar
Kuhn, S. L. & Zwyns, N.. 2014. Rethinking the initial Upper Paleolithic. Quaternary International 347: 2938.Google Scholar
Kuzmin, Y. V. & Keates, S. G.. 2018. Siberia and neighbouring regions in the Last Glacial Maximum: did people occupy northern Eurasia at that time? Archaeological and Anthropological Sciences 10: 111–24.Google Scholar
Lacy, S. A. 2014. Oral Health and Its Implications in Late Pleistocene Western Eurasian Humans. PhD Dissertation, Washington University, Proquest/UMIDissertation Publishing.Google Scholar
Lahr, M. M. & Foley, R. A.. 2003. Demography, dispersal and human evolution in the last glacial period. In van Andel, T. H & Davies, W. (eds.), Neanderthals and Modern Humans in the European Landscape during the Last Glaciation: Archaeological Results of the Stage 3 Project. Cambridge: McDonald Institute Monographs, pp. 241–56.Google Scholar
Lalueza-Fox, C. & Gilbert, M. T. P.. 2011. Paleogenomics of archaic hominins. Current Biology 21: R1002–9.Google Scholar
Lalueza-Fox, C., Sampietro, M. L., Caramelli, D. et al. 2005. Neandertal evolutionary genetics: mitochondrial DNA data from the Iberian Peninsula. Molecular Biology and Evolution 22 (4): 1077–81.Google Scholar
Lalueza-Fox, C., Rosas, A., Estalrrich, A. et al. 2011. Genetic evidence for patrilocal mating behaviour among Neandertal groups. Proceedings of the National Academy of Sciences USA 108 (1): 250–3.Google Scholar
Lambeck, K., Yokoyama, Y. & Purcell, T.. 2002. Into and out of the Last Glacial Maximum: sea-level change during Oxygen Isotope Stages 3 and 2. Quaternary Science Reviews 21: 343–60.Google Scholar
Lande, R. 1993. Risks of population extinction from demographic and environmental stochasticity and random catastrophes. The American Naturalist 142 (6): 911–27.Google Scholar
Langlais, M., Costamagno, S., Laroulandie, V. et al. 2012. The evolution of Magdalenian societies in South-West France between 18,000 and 14,000 calBP: changing environments, changing tool kits. Quaternary International 272–3: 138–49.Google Scholar
Langley, M. C. 2015. Investigating maintenance and discard behaviours for osseous projectile points: a Middle to Late Magdalenian (c. 19,000–14,000 cal. BP) example. Journal of Anthropological Archaeology 40: 340–60.Google Scholar
Langley, M. C. 2018. Magdalenian children: projectile points, portable art and playthings. Oxford Journal of Archaeology 37 (1): 324.Google Scholar
Langley, M. C. 2020. Space to play: identifying children’s sites in the Pleistocene archaeological record. Evolutionary Human Sciences 2: e41, http://doi.org/10.1017/ehs.2020.29.Google Scholar
Langley, M. C. & Litster, M. 2018. Is it ritual? Or is it children? Distinguishing consequences of play from ritual actions in the prehistoric archaeological record. Current Anthropology 59 (5): 616–43.Google Scholar
Langley, M. C & Street, M.. 2013. Long range inland-coastal networks during the Late Magdalenian: evidence for individual acquisition of marine resources at Andernach-Martinsberg, German Central Rhineland. Journal of Human Evolution 64: 457–65.Google Scholar
Larsson, L. 2000. Plenty of mammoths but no humans? Scandinavia during the Middle Weichselian. In Roebroeks, W, Mussi, M., Svoboda, J. & Fennema, K. (eds.), Hunters of the Golden Age. The Mid Upper Palaeolithic of Eurasia 30,000–20,000 BP. Leiden: University of Leiden Press, pp. 155–63.Google Scholar
Layton, R. & O’Hara, S.. 2010. Human social evolution: a comparison of hunter-gatherer and chimpanzee social organisation. Proceedings of the British Academy 158: 83113.Google Scholar
Layton, R., O’Hara, S. & Bilsborough, A.. 2012. Antiquity and social functions of multilevel social organization among human hunter-gatherers. International Journal of Primatology 33: 1215–45.Google Scholar
Lee, R. B. 1979. The !Kung San: Men, Women and Work in a Foraging Society. Cambridge: Cambridge University Press.Google Scholar
Lee, R. B. & DeVore, I.. 1968. Man the Hunter. New York: Aldine.Google Scholar
Leesch, D & Bullinger, J.. 2012. Identifying dwellings in Upper Palaeolithic open-air sites: the Magdalenian site at Monruz and its contribution to analysing palimpsests. In Niekus, M. J. L. T., Barton, R. N. E., Street, M. & Terberger, Th (eds.), A Mind Set on Flint. Studies in Honour of Dick Stapert. Barkhuis, Groningen University Library, Groningen, pp. 165–81.Google Scholar
Leesch, D., Müller, W., Nielsen, E. & Bullinger, J.. 2012. The Magdalenian in Switzerland: re-colonization of a newly accessible landscape. Quaternary International 272–3: 191208.Google Scholar
Leonardi, M., Librado, P., Der Sarkissian, C. et al. 2017. Evolutionary patterns and processes: lessons from ancient DNA. Systematic Biology 66 (1): e1e29.Google Scholar
Leroi-Gourhan, A. 1968. The Art of Prehistoric Man in Western Europe. London: Thames and Hudson.Google Scholar
Leroy, S. A. G., Arpe, K. & Mikolajewicz, U.. 2011. Vegetation context and climatic limits of the Early Pleistocene hominin dispersal in Europe. Quaternary Science Reviews 30: 1448–63.Google Scholar
Lesnik, J. J. & Sams, A. J.. 2014. Using resampling statistics to test male interment bias: applications for looted and commingled prehistoric remains in Peru and reassessment of Neandertal burials. PaleoAnthropology 2014: 463–9.Google Scholar
Lewis, S. G., Ashton, N. & Jacobi, R.. 2011. Testing human presence during the Last Interglacial (MIS 5e): a review of British evidence. In Ashton, N., Lewis, S. G. & Stringer, C. (eds.), The Ancient Human Occupation of Britain. Developments in Quaternary Science 14, pp. 125164, www.sciencedirect.com/bookseries/developments-in-quaternary-sciences/vol/14/suppl/CGoogle Scholar
Li, F., Kuhn, S. L., Chen, F. et al. 2018. The easternmost Middle Paleolithic (Mousterian) from Jinsitai Cave, north China. Journal of Human Evolution 114: 7684.Google Scholar
Lisiecki, L. E. & Raymo, M. E.. 2005. A Pliocene‐Pleistocene stack of 57 globally distributed benthic δ18O records. Paleoceanography 20 (1). http://doi.org//10.1029/2004PA001071.Google Scholar
Liu, W., Martinón-Torres, M., Cai, Y-J, et al. 2015. The earliest unequivocally modern humans in southern China. Nature 526: 696700.Google Scholar
Livi-Bacci, M. 2015. What we can and cannot learn from the history of world population. Population Studies 69 (Supplement 1): S21S28.Google Scholar
Livi-Bacci, M. 2017. A Concise History of World Population (6th ed.). Malden, MA: Blackwell.Google Scholar
Llamas, B., Willerslev, E. & Orlando, L.. 2016. Human evolution: a tale from ancient genomes. Philosophical Transactions of the Royal Society B 372: 20150484.Google Scholar
Llamas, B., Valverde, G., Fehren-Schmitz, L., Weyrich, L. S., Cooper, A. & Haak, W.. 2017. From the field to the laboratory: controlling DNA contamination in human ancient DNA research in the high-throughput sequencing era. STAR: Science and Technology of Archaeological Research 3 (1): 114.Google Scholar
Locht, J-L., Hérisson, D., Goval, E. et al. 2016. Timescales, space and culture during the Middle Palaeolithic in northwestern France. Quaternary International 411: 129–48.Google Scholar
Loog, L., Mirazón Lahr, M., Kovacevic, M., Manica, A., Eriksson, A. & Thomas, M. G.. 2017. Estimating mobility using sparse data: application to human genetic variation. Proceedings of the National Academy of Sciences USA 114 (46): 12213–18.Google Scholar
Lordkipanidze, D., Vekua, A., Ferring, R. et al. 2005. The earliest toothless hominin skull. Nature 434 (7034): 717–18.Google Scholar
Lordkipanidze, D., Jashashvili, T., Vekua, A. et al. 2007. Postcranial evidence from early Homo from Dmanisi, Georgia. Nature 449: 305–10.Google Scholar
Lordkipanidze, D., Ponce de Léon, M. S, Margvelashvili, A. et al. 2013. A complete skull from Dmanisi, Georgia, and the evolutionary biology of early Homo. Science 342: 326–31.Google Scholar
Low, B. S. 1994. Men in the demographic transition. Human Nature 5 (3): 223–53.Google Scholar
Lowe, J., Barton, N., Blockley, S. et al. 2012. Volcanic ash layers illuminate the resilience of Neanderthals and early modern humans to natural hazards. Proceedings of the National Academy of Sciences USA 109 (34): 13532–7.Google Scholar
Lozano, M., Mosquera, M., Bermúdez de Castro, J. M, Arsuaga, J. L & Carbonell, E. 2009. Right handedness of Homo heidelbergensis from Sima de los Huesos (Atapuerca, Spain) 500,000 years ago. Evolution and Human Behavior 30 (5): 369–76.Google Scholar
Ludwig, L., Pinto, J. G., Raible, C. C. & Shao, Y.. 2017. Impacts of surface boundary conditions on regional climate model simulations of European climate during the Last Glacial Maximum. Geophysical Research Letters 44: 5086–95.Google Scholar
Ludwig, P., Shao, Y., Kehl, M. & Weniger, G. C.. 2018. The Last Glacial Maximum and Heinrich event I on the Iberian Peninsula: a regional climate modelling study for understanding human settlement patterns. Global and Planetary Change 170: 3447.Google Scholar
Lugli, F., Cipriani, A., Arnaud, J., Arzarello, M., Peretto, C. & Benazzi, S.. 2017. Suspected limited mobility of a Middle Pleistocene woman from Southern Italy: strontium isotopes of a human deciduous tooth. Scientific Reports: 8615.Google Scholar
Lugli, F., Cipriani, A., Capecchi, G. et al. 2019. Strontium and stable isotope evidence of human mobility strategies across the Last Glacial Maximum in southern Italy. Nature Ecology & Evolution 3 (6): 905–11.Google Scholar
Lycett, S. J. 2015. Cultural evolutionary approaches to artifact variation over time and space: basis, progress, and prospects. Journal of Archaeological Science 56: 2131.Google Scholar
Lycett, S. J. & Gowlett, J. A. J.. 2008. On questions surrounding the Acheulean ‘tradition’. World Archaeology 40 (3): 295315.Google Scholar
Lycett, S. J. & Norton, C. J.. 2010. A demographic model for Palaeolithic technological evolution: the case of East Asia and the Movius Line. Quaternary International 211: 5565.Google Scholar
Lycett, S. J. & von Cramon-Taubadel, N.. 2008. Acheulean variability and hominin dispersals: a model-bound approach. Journal of Archaeological Science 35: 553–62.Google Scholar
Lyons, E. J., Amos, W., Berkley, J. A. et al. 2009. Homozygosity and risk of childhood death due to invasive bacterial disease. BMC Medical Genetics 10 (1): 55.Google Scholar
MacDonald, K. 2018. Fire-free hominin strategies for coping with cool winter temperatures in North-Western Europe from before 800,000 to circa 400,000 years ago. PaleoAnthropology 2018: 726.Google Scholar
MacDonald, K., Martinón-Torres, M., Dennell, R. W. & Bermúdez de Castro, J. M. 2012. Discontinuity in the record for hominin occupation in south-western Europe: implications for occupation of the middle latitudes of Europe. Quaternary International 271: 8497.Google Scholar
Mace, G. M., Collar, N. J., Gaston, et al. 2008. Quantification of extinction risk: IUCN’s system for classifying threatened species. Conservation Biology: 1424–42.Google Scholar
Madurell-Malapeira, J., Alba, D. M., Minwer-Barakat, R., Aurell-Garrido, J. & Moyà-Solà, S.. 2012. Early human dispersals into the Iberian Peninsula: a comment on Martínez et al. (2010) and Garcia et al. (2011). Journal of Human Evolution 62: 169–73.Google Scholar
Magyari, E. K., Veres, D., Wennrich, V. et al. 2014. Vegetation and environmental responses to climate forcing during the Last Glacial Maximum and deglaciation in the East Carpathians: attenuated response to maximum cooling and increased biomass burning. Quaternary Science Reviews 106: 278–94.Google Scholar
Maher, L. A & Conkey, M.. 2019. Homes for hunters? Exploring the concept of home at hunter-gatherer sites in Upper Palaeolithic Europe and Epipalaeolithic Southwest Asia. Current Anthropology 60 (1): 91137.Google Scholar
Maier, A. 2012. Regional groups and social interaction during the Central European Magdalenian. Notae Praehistoricae 32: 121–32.Google Scholar
Maier, A. 2015. The Central European Magdalenian: Regional Diversity and Internal Variability. Dordrecht: Springer.Google Scholar
Maier, A. 2017. Population and settlement dynamics from the Gravettian to the Magdalenian. Mitteilungen der Gesellschaft für Urgeschichte 26: 83101.Google Scholar
Maier, A. & Zimmermann, A.. 2017. Populations headed south? The Gravettian from a palaeodemographic point of view. Antiquity 91: 573–88.Google Scholar
Maier, A., Lehmkuhl, F., Ludwig, P. et al. 2016. Demographic estimates of hunter-gatherers during the Last Glacial Maximum in Europe against the background of palaeoenvironmental data. Quaternary International 425: 4961.Google Scholar
Maier, A., Liebermann, C. & Pfeifer, S. J.. 2020. Beyond the Alps and Tatra Mountains—the 20–14 ka repopulation of the northern mid-latitudes as inferred from palimpsests deciphered with keys from western and central Europe. Journal of Paleolithic Archaeology (3): 398-452.Google Scholar
Mallegni, F. & Fabbri, F.. 1995. The human skeletal remains from the Upper Palaeolithic burials found in Romito cave (Papasidero, Cosenza, Italy). Bulletins de Mémoires de la Société d’Anthropologie de Paris 7 (3–4): 99137.Google Scholar
Malthus, T. R. 1872 [1973]. An Essay on the Principle of Population (7th ed.). London: J.M. Dent & Sons.Google Scholar
Mance, M. J. 2008. Keeping infants warm: challenges of hypothermia. Advances in Neonatal Care 8 (1): 612.Google Scholar
Mania, D. 1991. The zonal division of the Lower Palaeolithic open-air site at Bilzingsleben. Anthropologie 29: 1724.Google Scholar
Mania, D. & Mania, U.. 2005. The natural and socio-cultural environment of Homo erectus at Bilzingsleben, Germany. In Gamble, C. & Porr, M. (eds.), The Hominid Individual in Context: Archaeological Investigations of Lower and Middle Palaeolithic Landscapes, Locales and Artefacts. London: Routledge, pp. 98114.Google Scholar
Marín-Arroyo, A. B. M. 2009. Economic adaptations during the Late Glacial in northern Spain: a simulation approach. Before Farming 2009 (2): 118.Google Scholar
Marlowe, F. 2001. Male contribution to diet and female reproductive success among foragers. Current Anthropology 42 (5): 755759.Google Scholar
Marlowe, F. W. 2005. Hunter-gatherers and human evolution. Evolutionary Anthropology 14: 5467.Google Scholar
Marlowe, F. W. 2010. The Hadza: Hunter-Gatherers of Tanzania. Berkeley: University of California Press.Google Scholar
Marom, A., McCullagh, J. S., Higham, T. F. G., Sinitsyn, A. A. & Hedges, R. E. M.. 2012. Single amino acid radiocarbon dating of Upper Paleolithic modern humans. Proceedings of the National Academy of Sciences USA 109 (18): 6878–81.Google Scholar
Marom, A., McCullagh, J. S. O, Higham, T. F. G. & Hedges, R. E. M.. 2013. Hydroxyproline dating: experiments on the 14 C analysis of contaminated and low-collagen bones. Radiocarbon 55 (2): 698708.Google Scholar
Marreiros, J. & Bicho, N.. 2013. Lithic technology variability and human ecodynamics during the Early Gravettian of Southern Iberian Peninsula. Quaternary International 318: 90101.Google Scholar
Martínez, I., Rosa, M., Quam, R. et al. 2013. Communicative capacities in Middle Pleistocene humans from the Sierra de Atapuerca in Spain. Quaternary International 295: 94101.Google Scholar
Martínez, K. & Garriga, J. G.. 2016. On the origin of the European Acheulian. Journal of Anthropological Archaeology 44: 87104.Google Scholar
Martínez, K., Garcia, J., Carbonell, E. et al. 2010. A new lower Pleistocene archaeological site in Europe (Vallparadís, Barcelona, Spain). Proceedings of the National Academy of Sciences USA 107 (13): 5762–7.Google Scholar
Martinón-Torres, M., Bermúdez de Castro, J. M, Gómez-Robles, A. et al. 2007. Dental evidence on the hominin dispersals during the Pleistocene. Proceedings of the National Academy of Sciences USA 104 (33): 13279–82.Google Scholar
Maslin, M. A. & Brierley, C. M.. 2015. The role of orbital forcing in the Early Middle Pleistocene Transition. Quaternary International 389: 4755.Google Scholar
Matisoo-Smith, E. & Horsburgh, K. A.. 2012. DNA for Archaeologists. Walnut Creek, CA: Left Coast Press.Google Scholar
Mayr, E. 1964. Systematics and the Origin of Species: From the Viewpoint of a Zoologist. Dover: Dover Publications.Google Scholar
Mazet, O., Rodríguez, W., Grusea, S., Boitard, S. & Chikhi, L. 2016. On the importance of being structured: instantaneous coalescence rates and human evolution—lessons for ancestral population size inference? Heredity 116 (4): 362–71.Google Scholar
McBrearty, S. & Brooks, A. S.. 2000. The revolution that wasn’t: a new interpretation of the origin of modern human behaviour. Journal of Human Evolution 39 (5): 453–63.Google Scholar
McCoid, C. H. & McDermott, L. D.. 1996. Towards decolonizing gender: female vision in the Upper Paleolithic. American Anthropologist 98 (2): 319–26.Google Scholar
McFadden, C. & Oxenham, M. F.. 2017. The D0-14/D ratio: a new paleodemographic index and equation for estimating total fertility rates. American Journal of Physical Anthropology 165 (3): 471–9.Google Scholar
McLaughlin, T. R. 2019. On applications of space-time modelling with open-source 14C age calibration. Journal of Archaeological Method and Theory 26: 479501.Google Scholar
McNabb, J. 2017. Journeys in space and time: assessing the link between Acheulean handaxes and genetic explanations. Journal of Archaeological Science: Reports 13: 403–14.Google Scholar
Meindl, R. S. & Russell, K. F.. 1998. Recent advances in method and theory in paleodemography. Annual Review of Anthropology 27: 375–99.Google Scholar
Melchionna, M., Di Febbraro, M., Carotenuto, F. et al. 2018. Fragmentation of Neanderthals’ pre-extinction distribution by climate change. Palaegeography, Palaeoclimatology, Palaeoecology 496: 146–54.Google Scholar
Mellars, P. 1996. The Neanderthal Legacy: An Archaeological Perspective from Western Europe. Princeton, NJ: Princeton University Press.Google Scholar
Mellars, P. 2005. The impossible coincidence: a single‐species model for the origins of modern human behavior in Europe. Evolutionary Anthropology 14 (1): 1227.Google Scholar
Mellars, P. 2006. Archeology and the dispersal of modern humans in Europe: Deconstructing the ‘Aurignacian’. Evolutionary Anthropology 15 (5): 167–82.Google Scholar
Mellars, P. 2010. Neanderthal symbolism and ornament manufacture: the bursting of a bubble. Proceedings of the National Academy of Sciences USA 107 (47): 20147–8.Google Scholar
Mellars, P. & French, J. C.. 2011. Tenfold population increase in Western Europe at the Neanderthal-to-modern human transition. Science 333: 623–7.Google Scholar
Mellars, P. & Stringer, C.. 1989. The Human Revolution: Behavioural and Biological Perspectives on the Origins of Modern Humans. Edinburgh: Edinburgh University Press.Google Scholar
Méndez-Quintas, E., Santonja, M., Pérez-González, A., Duval, M., Demuro, M. & Arnold, L. J.. 2018. First evidence of an extensive Acheulean large cutting tool accumulation in Europe from Porto Maior (Galicia, Spain). Scientific Reports 8: 3082.Google Scholar
Metcalf, C. J. E. & Pavard, S.. 2006. Why evolutionary biologists should be demographers. Trends in Ecology and Evolution 22 (4): 205–12.Google Scholar
Meyer, M., Kircher, M., Gansauge, M. T. et al. 2012. A high-coverage genome sequence from an archaic Denisovan individual. Science 338 (6104): 222–6.Google Scholar
Meyer, M., Arsuaga, J-L, de Filippo, C. et al. 2016. Nuclear DNA sequences from the Middle Pleistocene Sima de los Huesos hominins. Nature 531: 504–7.Google Scholar
Mgeladze, A., Lordkipanidze, D., Moncel, M-H et al. 2008. Hominin occupations at the Dmanisi site, Georgia, Southern Caucasus: raw materials and technical behaviours of Europe’s first hominins. Journal of Human Evolution 60: 571–96.Google Scholar
Migliano, A. B., Vinicius, L. & Lahr, M. M.. 2007. Life history trade-offs explain the evolution of human pygmies. Proceedings of the National Academy of Sciences USA 104 (51): 20216–19.Google Scholar
Migliano, A. B., Page, A. E., Gómez-Gardeñes, J. et al. 2017. Characterization of hunter-gatherer networks and implications for cumulative culture. Nature Human Behaviour 1 (2): 16.Google Scholar
Migliano, A. B., Battison, F., Viguier, S. et al. 2020. Hunter-gatherer multilevel sociality accelerates cumulative cultural evolution. Science Advances 6 (9): eaax5913.Google Scholar
Milks, A., Parker, D. & Pope, M.. 2019. External ballistics of Pleistocene hand-thrown spears: experimental performance data and implications for human evolution. Scientific Reports 9: 820.Google Scholar
Miller, R. 2012. Mapping the expansion of the Northwest Magdalenian. Quaternary International 272: 209–30.Google Scholar
Mithen, S. 1996. The Prehistory of the Mind: A Search for the Origins of Art, Religion and Science. London: Thames and Hudson.Google Scholar
Mix, A., C., Bard, E. & Schneider, R.. 2001. Environmental processes of the ice age: land, oceans, glaciers (EPILOG). Quaternary Science Reviews 20 (4): 627–57.Google Scholar
Modesto-Mata, M., Dean, M. C., Lacruz, R. S. et al. 2020. Short and long period growth markers of enamel formation distinguish European Pleistocene hominins. Scientific Reports 10: 4665.Google Scholar
Moncel, M-H. & Schreve, D.. 2016. The Acheulean in Europe: origins, evolution, and dispersal. Quaternary International 411: 18.Google Scholar
Moncel, M-H., Despriée, J., Voinchet, P. et al. 2013. Early evidence of Acheulean settlement in Northwestern Europe—La Noira site, a 700 000-year-old occupation in the center of France. PloS 8 (11): e75529.Google Scholar
Moncel, M-H., Santagata, C., Pereira, A. et al. 2020. The origin of early Acheulean expansion in Europe 700 ka ago: new findings at Notarchirico (Italy). Scientific Reports 10: 13802.Google Scholar
Monge, J. & Mann, A.. 2007. Paleodemography of extinct hominin populations. In Henke, W & Tattersall, I (eds.), Handbook of Paleoanthropology. Berlin: Springer, pp. 673700.Google Scholar
Morabia, A, 2008. The relation of childhood mortality to mobility in contemporary foragers. Human Ecology 36: 931–2.Google Scholar
Morales, T. M. 1987. An examination of infanticide practices among mobile and sedentary hunter-gatherers. Haliksa’i: University of New Mexico contributions to Anthropology VI: 119.Google Scholar
Moreau, L., Brandl, M. & Nigst, P. R.. 2016. Did prehistoric foragers behave in an economically irrational manner? Raw material availability and technological organisation at the early Gravettian site of Willendorf II (Austria). Quaternary International 406: 8494.Google Scholar
Morgan, S. P. & Hagewen, K. J.. 2005. Fertility. In Poston, D. L. & Micklin, M. (eds.), Handbook of Population. Dordrecht: Springer, pp. 229–50.Google Scholar
Morin, E. 2008. Evidence for declines in human population densities during the early Upper Palaeolithic in western Europe. Proceedings of the National Academy of Sciences USA 105 (1): 4853.Google Scholar
Moroni, A., Boscato, P. & Ronchitelli, A.. 2013. What roots for the Uluzzian? Modern behaviour in Central-Southern Italy and hypotheses on AMH dispersal routes. Quaternary International 316: 2744.Google Scholar
Mosquera, M., Ollé, A. & Rodríguez, X. P.. 2013. From Atapuerca to Europe: tracing the earliest peopling of Europe. Quaternary International 295: 130–7.Google Scholar
Mosquera, M., Saladié, P., Ollé, A. et al. 2015. Barranc de la Boella (Catalonia, Spain): an Acheulean elephant butchering site from the European late Early Pleistocene. Journal of Quaternary Science 30 (7): 651–66.Google Scholar
Mourier, T., Ho, S. Y. W., Gilbert, M. T. P., Willerslev, E. & Orlando, L.. 2012. Statistical guidelines for detecting past population shifts using ancient DNA. Molecular Biology and Evolution 29 (9): 2241–51.Google Scholar
Movius, H. L. 1948. The Lower Palaeolithic cultures of southern and eastern Asia. Transactions of the American Philosophical Society 38: 329426.Google Scholar
Müller, J. & Diachenko, A., A. 2019. Tracing long-term demographic changes: The issue of spatial scales. PloS one 14 (1): e0208739.Google Scholar
Müller, U. C., Pross, J., Tzedakis, P. C. et al. 2011. The role of climate in the spread of modern humans into Europe. Quaternary Science Reviews 2011: 273–9.Google Scholar
Münzel, S. C. & Conard, N. J.. 2004. Cave bear hunting in the Hohle Fels, a cave site in the Ach Valley, Swabian Jura. Revue de Paléobiologie 23 (2): 877–85.Google Scholar
Musil, R. 2011. Gravettian environmental changes in a N-S transect of central Europe. Central European Journal of Geosciences 3 (2): 147–54.Google Scholar
Mussi, M. 1999. The Neanderthals in Italy: a tale of many caves. In Roebroeks, W. & Gamble, C. (eds.), The Middle Palaeolithic Occupation of Europe. Leiden: University of Leiden Press, pp. 4980.Google Scholar
Mussi, M. 2007. Women of the middle latitudes: the earliest peopling of Europe from a female perspective. In Roebroeks, W. (ed.), Guts and Brains: An Integrative Approach to the Hominin Record. Leiden: Leiden University Press, pp. 165–83.Google Scholar
Mussi, M. 2015. Encoding and decoding the message: the case of mid Upper Palaeoltihic female imagery. In Coward, F., Hosfield, R., Pope, M. & Wenban-Smith, F. (eds.), Settlement, Society and Cognition in Human Evolution: Landscapes in Mind. Cambridge: Cambridge University Press, pp. 275–87.Google Scholar
Mussi, M., Roebroeks, W. & Svoboda, J.. 2000. Hunters of the Golden Age: an introduction. In Roebroeks, W., Mussi, M., Svoboda, J. & Fennema, K. (eds.), Hunters of the Golden Age: The Mid Upper Palaeolithic of Eurasia 30,000–20,000 BP. Leiden: University of Leiden Press, pp. 111.Google Scholar
Muttoni, G., Scardia, G. & Kent, D. V.. 2010. Human migration into Europe during the late Early Pleistocene climate transition. Palaeogeography, Palaeoclimatology, Palaeoecology 296: 7993.Google Scholar
Muttoni, G., Scardia, G., Kent, D. V. et al. 2011. First dated human occupation of Italy at ~0.85 ma during the late Early Pleistocene climatic transition. Earth and Planetary Science Letters 307: 241–52.Google Scholar
Muttoni, G., Scardia, G. & Kent, D. V .. 2013. A critique of evidence for human occupation of Europe older than the Jaramillo subchron (~1 ma): comment on ‘The oldest human fossil in Europe from Orce, (Spain)’ by Toro-Moyano et al. 2013. Journal of Human Evolution 65: 746–9.Google Scholar
Muttoni, G., Kent, D. V., Scardina, G. & Martin, R. A.. 2015. Bottleneck at Jaramillo for human migration to Iberia and the rest of Europe? Journal of Human Evolution 80: 187–90.Google Scholar
Muttoni, G., Sirakov, N., Guadelli, J-L et al. 2017. An early Brunhes (<0.78 ma) age for the Lower Paleolithic tool-bearing Kozarnika cave sediments, Bulgaria. Quaternary Science Reviews 178: 113.Google Scholar
Muttoni, G., Scardia, G. & Kent, D. V.. 2018. Early hominins in Europe: the Galerian migration hypothesis. Quaternary Science Reviews 180: 129.Google Scholar
Nakahashi, W. 2017. The effect of trauma on Neanderthal culture: a mathematical analysis. HOMO Journal of Comparative Human Biology 68: 83100.Google Scholar
Nakahashi, W. & Feldman, M. W.. 2014. Evolution of division of labor: emergence of different activities among group members. Journal of Theoretical Biology 348: 6579.Google Scholar
Nakahashi, W., Horiuchi, S. & Ihara, Y.. 2018. Estimating hominid life history: the critical interbirth interval. Population Ecology 60: 127–42Google Scholar
Nalawade-Chavan, S., McCullagh, J. & Hedges, R.. 2014. New hydroxyproline radiocarbon dates from Sungir, Russia, confirm Early Mid Upper Palaeolithic burials in Eurasia. PloS One 9 (1): e76896.Google Scholar
Nava, A., Coppa, A., Coppola, D. et al. 2017. Virtual histological assessment of the prenatal life history and age at death of the Upper Paleolithic fetus from Ostuni (Italy). Scientific Reports 7: 9427.Google Scholar
Nava, A., Lugli, F., Romandini, M. et al. 2020. Early life of Neanderthals. Proceedings of the National Academy of Sciences USA 117 (46): 28719–26.Google Scholar
Neruda, P. & Nerudová, Z.. 2013. The Middle-Upper Palaeolithic transition in Moravia in the context of the Middle Danube region. Quaternary International 294: 319.Google Scholar
Nerudová, Z. & Neruda, P.. 2015. Moravia between Gravettian and Magdalenian. In Sázelová, S., Novák, M. & Mizerová, A. (eds.), Forgotten Times and Spaces. New Perspectives in Paleoanthropological, Paleothnological and Archaeological Studies. Institute of Archaeology of the Czech Academy of Sciences, Masaryk University: Brno, pp. 378–94.Google Scholar
Nerudová, Z., Doláková, N. & Novák, J.. 2016. New information augmenting the picture of local environment at the LGM/LGT in the context of the Middle Danube region. The Holocene 26 (9): 1345–54.Google Scholar
Nerudová, Z., Vaníčková, V., Tvrdý, Z., Ramba, J., Bílek, O. & Kostrhun, P.. 2019. The woman from the Dolní Věstonice 3 burial: a new view of the face using modern technologies. Archaeological and Anthropological Sciences 11: 2527–38.Google Scholar
Nettle, D., Gibson, M. A., Lawson, D. W. & Sear, R.. 2013. Human behavioral ecology: current research and future prospects. Behavioral Ecology 24 (5): 1031–40.Google Scholar
Nicholson, C. M. 2017. Eemian paleoclimate zones and Neanderthal landscape-use: a GIS model of settlement patterning during the last interglacial. Quaternary International 438 (B): 144–57.Google Scholar
Nicholson, C. M. 2020. Shifts along a spectrum: a longitudinal study of the western Eurasian realized climate niche. Environmental Archaeology 25 (4): 381–96.Google Scholar
Nielsen, T. K., Benito, B. M., Svenning, J-C et al. 2017. Investigating Neanderthal dispersal above 55°N during the Last Interglacial Complex. Quaternary International 431: 88103.Google Scholar
Nigst, P. R. 2012. The Early Upper Palaeolithic of the Middle Danube Region. Leiden: Leiden University Press.Google Scholar
Nigst, P. R., Haesaerts, P., Damblon, F. et al. 2014. Early modern human settlement of Europe north of the Alps occurred 43,500 years ago in a cold steppe-type environment. Proceedings of the National Academy of Sciences USA 111 (40): 14394–9.Google Scholar
Niskanen, M., Ruff, C. B., Holt, B., Sládek, V. & Berner, M.. 2018. Temporal and geographic variation in body size and shape of Europeans from the Late Pleistocene to recent times. In Ruff, C. B (eds.), Skeletal Variation and Adaptations in Europeans: Upper Paleolithic to the Twentieth Century. Hoboken, NJ: John Wiley & Sons, pp. 4989.Google Scholar
Nixon, K. C. & Wheeler, Q. D.. 1990. An amplification of the phylogenetic species concept. Cladistics 6: 211–23.Google Scholar
Noiret, P. 2009. Le Paléolithique Supérieur de Moldavie: Essai de Synthèse d’une Évolution Multi-Culturelle. Liège: Études et Recherches Archéologiques de l’Université de Liège (ERAUL) 121.Google Scholar
Nowell, A. 2010a. Working memory and the speed of life. Current Anthropology 51 (Supplement 1): S121S133.Google Scholar
Nowell, A. 2010b. Defining behavioural modernity in the context of Neandertal and anatomically modern populations. Annual Review of Anthropology 39: 437–52.Google Scholar
Nowell, A. 2020. Reconsidering the personhood of Gravettian infants. Journal of Anthropological Research 76 (2): 232–50.Google Scholar
Nowell, A. & Chang, M. L.. 2014. Science, the media and interpretations of Upper Paleolithic figurines. American Anthropologist 116 (3): 562–77.Google Scholar
Nowell, A. & French, J. C.. 2020. Adolescence and innovation in the European Upper Palaeolithic. Evolutionary Human Sciences 2: e36.Google Scholar
Nowell, A. & White, M.. 2010. Growing up in the Middle Pleistocene. Life history strategies and their relationship to Acheulian industries. In Nowell, A & Davidson, I (eds.), Stone Tools and the Evolution of Human Cognition. Boulder (CO): University of Colorado Press, pp. 6781.Google Scholar
O’Brien, M. J. & Buchanan, B.. 2017. Cultural learning and the Clovis colonization of North America. Evolutionary Anthropology 26: 270–84.Google Scholar
O’Connell, J. F., Hawkes, K. & Blurton Jones, N. G.. 1999. Grandmothering and the evolution of Homo erectus. Journal of Human Evolution 36: 461–85.Google Scholar
Oakley, K. P., Andrews, P., Keeley, L. H. & Clark, J. D.. 1977. A re-appraisal of the Clacton spearpoint. Proceedings of the Prehistoric Society 43: 1330.Google Scholar
Ogilvie, M. D., Curran, B. K. & Trinkaus, E.. 1989. Incidence and patterning of dental enamel hypoplasia among the Neandertals. American Journal of Physical Anthropology 79: 2541.Google Scholar
Oliva, M. 1988. Discovery of a Gravettian mammoth bone hut at Milovice (Moravia, Czechoslovakia). Journal of Human Evolution 17: 787–90.Google Scholar
Olive, M., Pigeot, N. & Bignon-Lau, O.. 2019. Un campement magdalénien à Étiolles (Essonne). Des activités à la microsociologie d’un habitat. Gallia Préhistoire (59): 47108.Google Scholar
Ollé, A., Mosquera, M., Rodríguez-Álvarez, X. P. et al. 2016. The Acheulean from Atapuerca: three steps forward, one step back. Quaternary International 411: 316–28.Google Scholar
Oosterbeek, L., Grimaldi, S., Rosinia, P., Cura, S., Cumha, P. P. & Martins, A.. 2010. The earliest Pleistocene archaeological sites in western Iberia: present evidence and research prospects. Quaternary International 223–4: 399407.Google Scholar
Orlando, L. & Cooper, A.. 2014. Using ancient DNA to understand evolutionary and ecological processes. Annual Review of Ecology, Evolution, and Systematics 45: 573–98.Google Scholar
Orschiedt, J. 2008. Der Fall Krapina: Neue Ergebnisse zur Frage von Kannibalismus beim Neandertaler. Quartär 55: 6381.Google Scholar
Orschiedt, J. 2013. Bodies, bits and pieces: burials from the Magdalenian and the Late Palaeolithic. In Pastoors, A & Auffermann, B. (eds.), Pleistocene Foragers: Their Culture and Environment. Festschrift in Honour of Gerd-Christian Weniger for his Sixtieth Birthday. Mettmann: Wissenschaftliche Schriften des Neanderthal Museums, pp. 117132.Google Scholar
Orschiedt, J., Schüler, T., Połtowicz-Bobak, M., Bobak, D., Kozłowski, S. K. & Terberger, T.. 2017a. Human remains from Maszycka Cave (Woj. Małopoliskie/PL): The treatment of human bodies in the Magdalenian. Archäologisches Korrespondenzblatt 47: 423–38.Google Scholar
Orschiedt, J., Kierdorf, U., Schultz, M., Baales, M., von Berg, A. & Flohr, S.. 2017b. The Late Upper Palaeolithic human remains from Neuwied-Irlich, Germany. A rare find from the Late Glacial of Central Europe. Quartär 64: 203–16.Google Scholar
Osborn, A. J. & Hitchcock, R. K.. 2019. Information sharing in times of scarcity: drought strategies in the Kalahari Desert and the central plains of North America. In Lavi, N & Friesem, D. E. (eds.), Towards a Broader View of Hunter-Gatherer Sharing. Cambridge: MacDonald Institute Monographs, pp. 123–41.Google Scholar
Otte, M. 2015. Central Asia as a core area: Iran as an origin for the European Aurignacian. International Journal of the Society of Iranian Archaeologists 1 (1): 1923.Google Scholar
Otte, M. & Noiret, P. 2007. Le Gravettien du nord-ouest de l’Europe. PALEO. Revue d’Archéologie Préhistorique (19): 243–55.Google Scholar
Otterbein, K. 2011. The earliest evidence for warfare? A comment on Carbonell et al. Current Anthropology 52: 439.Google Scholar
Overmann, K. A. & Coolidge, F. L.. 2013. Human species and mating systems: Neandertal-Homo sapiens reproductive isolation and the archaeological and fossil records. Journal of Anthropological Sciences 91: 91110.Google Scholar
Oxillia, G., Peresani, M., Romandini, M. et al. 2015. Earliest evidence of dental caries manipulation in the Late Upper Palaeolithic. Scientific Reports 5: 12150.Google Scholar
Page, A. E. & French, J. C.. 2020. Reconstructing prehistoric demography: what role for extant hunter-gatherers? Evolutionary Anthropology 29 (6): 332–45.Google Scholar
Page, A. E., Viguier, S., Dyble, M. et al. 2016. Reproductive trade-offs in extant hunter-gatherers suggest adaptive mechanism for the Neolithic expansion. Proceedings of the National Academy of Sciences USA 113 (17): 4694–9.Google Scholar
Page, A. E., Chaudhary, N., Viguier, S. et al. 2017. Hunter-gatherer social networks and reproductive success. Scientific Reports 7 (1): 110.Google Scholar
Page, A. E., Myers, S., Dyble, M. & Migliano, A. B.. 2019. Why so many Agta boys? Explaining ‘extreme’ sex ratios in Philippine foragers. Evolutionary Human Sciences 1 e5, https://www.cambridge.org/core/journals/evolutionary-human-sciences/article/why-so-many-agta-boys-explaining-extreme-sex-ratios-in-philippine-foragers/83A0E976E1178011ACB6B21BEE1D2DF3Google Scholar
Pala, M., Olivieri, A., Achilli, A. et al. 2012. Mitochondrial DNA signals of Late Glacial recolonization of Europe from Near Eastern Refugia. American Journal of Human Genetics 90 (5): 915–24.Google Scholar
Palombo, M. R. 2016. Faunal dynamics during the Mid- Pleistocene revolution: what implications, if any, does this have for the appearance of the Acheulean in southwestern Europe? Quaternary International 411: 262–83.Google Scholar
Parés, J. M., Duval, M. & Arnold, L. J.. 2013. New views on an old move: hominin migration into Eurasia. Quaternary International 295: 512.Google Scholar
Parfitt, S. A. & Roberts, M. B.. 1999. Human modification of faunal remains. In Roberts, M. B & Parfitt, S. A. (eds.), Boxgrove: A Middle Pleistocene Hominid Site at Eartham Quarry, Boxgrove, West Sussex. London: English Heritage, pp. 398419.Google Scholar
Parfitt, S. A., Ashton, N. M., Lewis, S. G. et al. 2010. Early Pleistocene human occupation at the edge of the boreal zone in northwest Europe. Nature 466: 229–33.Google Scholar
Partiot, C., Trinkaus, E., Knüsel, C. J. & Villotte, S.. 2020. The Cro-Magnon babies: morphology and mortuary implications of the Cro-Magnon immature remains. Journal of Archaeological Science: Reports 30: 102257.Google Scholar
Patou-Mathis, M. 2000. Neanderthal subsistence behaviours in Europe. International Journal of Osteoarchaeology 10: 379–95.Google Scholar
Patton, H., Hubbard, A., Andreassen, K. et al. 2017. Deglaciation of the Eurasian ice sheet complex. Quaternary Science Reviews 169: 148–72.Google Scholar
Pavlov, P. & Indrelid, S.. 2000. Human occupation in Northeastern Europe during the period 35,000–18,000 BP. In Roebroeks, W, Mussi, M, Svoboda, J & Fennema, K (eds.), Hunters of the Golden Age. The Mid Upper Palaeolithic of Eurasia 30,000–20,000 BP. Leiden: University of Leiden Press, pp. 165172.Google Scholar
Pavlov, P., Svendsen, J. I. & Indrelid, S.. 2001. Human presence in the European Arctic nearly 40,000 years ago. Nature 413: 64–7.Google Scholar
Pearce, E. 2014. Modelling mechanisms of social network maintenance in hunter–gatherers. Journal of Archaeological Science 50: 403–13.Google Scholar
Pearce, E. 2018. Neanderthals and Homo sapiens: cognitively different kinds of humans? In Di Paolo, L. D., Di Vincenzo, F. & De Petrillo, F. (eds.), Evolution of Primate Social Cognition. Dordrecht: Springer, pp. 181–96.Google Scholar
Pearce, E. & Moutisou, T.. 2014. Using obsidian transfer distances to explore social network maintenance in late Pleistocene hunter-gatherers. Journal of Anthropological Archaeology 36: 1220.Google Scholar
Pearce, E., Stringer, C. & Dunbar, R. I.. 2013. New insights into differences in brain organization between Neanderthals and anatomically modern humans. Proceedings of the Royal Society of London B 280 (1758): 20130168.Google Scholar
Pearce, E., Shuttleworth, A., Grove, M. & Layton, R. 2014. The costs of being a high latitude hominin. In Dunbar, R. I. M., Gamble, C. & Gowlett, J. A. J. (eds.), Lucy to Language: The Benchmark Papers. Oxford: Oxford University Press, pp. 356–79.Google Scholar
Peeters, J. H. M. & Momber, G.. 2014. The southern North Sea and the human occupation of northwest Europe after the Last Glacial Maximum. Netherlands Journal of Geosciences 93 (1/2): 5570.Google Scholar
Pennington, R. 2001. Hunter-gatherer demography. In Panter-Brick, C., Layton, R. H. & Rowley-Conwy, P. (eds.), Hunter-Gatherers: An Interdisciplinary Perspective. Cambridge: Cambridge University Press, pp. 170204.Google Scholar
Peresani, M., Vanhaeren, M., Quaggiotto, E., Queffelec, A. & d’Errico, F. 2013. An ochered fossil marine shell from the Mousterian of Fumane Cave, Italy.PloS One 8 (7): e68752.Google Scholar
Peresani, M., Cristiani, E. & Romandini, M.. 2016. The Uluzzian technology of Grotta di Fumane and its implication for reconstructing cultural dynamics in the Middle-Upper Palaeolithic transition of Western Eurasia. Journal of Human Evolution 91: 3656.Google Scholar
Pesesse, D. 2017. Is it still appropriate to talk about the Gravettian? Data from lithic industries in Western Europe? Quartär 64: 107–28.Google Scholar
Petersen, W. 1975. A demographer’s view of prehistoric demography (and comments and replies). Current Anthropology 16 (2): 227–45.Google Scholar
Pétillon, J. M. 2016. Technological evolution of hunting implements among Pleistocene hunter–gatherers: Osseous projectile points in the middle and upper Magdalenian (19–14 ka cal BP). Quaternary International 414: 108–34.Google Scholar
Petr, M, Pääbo, S., Kelso, J. & Vernot, B.. 2019. Limits of long-term selection against Neanderthal introgression. Proceedings of the National Academy of Sciences USA 116 (5): 1639–44.Google Scholar
Petru, S. 2018. Identity and fear- burials in the Upper Palaeolithic. Documenta Praehistorica XLV: 613.Google Scholar
Pettitt, P. B. 1997. High resolution Neanderthals? Interpreting Middle Palaeolithic intrasite spatial data. World Archaeology 29 (2): 208–24.Google Scholar
Pettitt, P. 2011. The Palaeolithic Origins of Human Burial. London: Routledge.Google Scholar
Pettitt, P. 2017. Palaeolithic Western and North Central Europe. In Insoll, T (ed.), The Oxford Handbook of Prehistoric Figurines. Oxford: Oxford University Press, pp. 851–76.Google Scholar
Pettitt, P & White, M.. 2012. The British Palaeolithic: Human Societies at the Edge of the Pleistocene World. London: Routledge.Google Scholar
Peyrégne, S., Slon, V., Mafessoni, F. et al. 2019. Nuclear DNA from two early Neandertals reveals 80,000 years of genetic continuity in Europe. Science Advances 5 (6): eaaw5873.Google Scholar
Ponce de Léon, M. S., Golovanova, L., Doronichev, V. et al. 2008. Neanderthal brain size at birth provides insights into the evolution of human life history. Proceedings of the National Academy of Sciences USA 105 (37): 13764–8.Google Scholar
Pontzer, H., Raichlen, D. A., Wood, B. M., Mabulla, A. Z. P., Racette, S. B. & Marlowe, F. W.. 2012. Hunter-gatherer energetics and human obesity. PloS One 7 (7): e40503.Google Scholar
Porčić, M. 2015. Exploring the effects of assemblage accumulation on diversity and innovation rate estimates in neutral, conformist, and anti-conformist models of cultural transmission. Journal of Archaeological Method and Theory 22 (4): 1071–92.Google Scholar
Porčić, M. & Nikolić, M.. 2016. The Approximate Bayesian Computation approach to reconstructing population dynamics and size from settlement data: demography of the Mesolithic-Neolithic transition at Lepenski Vir. Archaeological and Anthropological Sciences 8 (1): 169–86.Google Scholar
Posth, C., Renaud, G., Mittnik, A. et al. 2016. Pleistocene mitochondrial genomes suggest a single major dispersal of non-Africans and a late glacial population turnover in Europe. Current Biology 26: 827–33.Google Scholar
Posth, C., Wissing, C., Kitagawa, K. et al. 2017. Deeply divergent archaic mitochondrial genome provides lower boundary for African gene flow into Neanderthals. Nature Communications 8: 16046.Google Scholar
Powell, A., Shennan, S. & Thomas, M. G. 2009. Late Pleistocene demography and the appearance of modern human behaviour. Science 324: 1298–301.Google Scholar
Power, C., Sommer, V. & Watts, I.. 2013. The seasonality thermostat: female reproductive synchrony and male behavior in monkeys, Neanderthals, and modern humans. PaleoAnthropology 2013: 3360.Google Scholar
Power, R. C. & Williams, F. L. E.. 2018. Evidence of increasing intensity of food processing during the Upper Paleolithic of Western Eurasia. Journal of Paleolithic Archaeology 1 (4): 281301.Google Scholar
Power, R. C., Salazar-García, D. C., Straus, L. G., González Morales, M. R. & Henry, A. G.. 2015. Microremains from El Mirón cave human dental calculus suggest a mixed plant-animal subsistence economy during the Magdalenian in Northern Iberia. Journal of Archaeological Science 60: 3946.Google Scholar
Poza-Rey, E. M., Lozano, M. and Arsuaga, J. L., 2017. Brain asymmetries and handedness in the specimens from the Sima de los Huesos site (Atapuerca, Spain). Quaternary International 433: 3244.Google Scholar
Poza-Rey, E. M., Gómez-Robles, A. & Arsuaga, J. L.. 2019. Brain size and organization in the Middle Pleistocene hominins from Sima de los Huesos. Inferences from endocranial variation. Journal of Human Evolution 129: 6790.Google Scholar
Prat, S., Péan, S. C., Crépin, L. et al. 2011. The oldest anatomically modern humans from far southeast Europe: direct dating, culture and behavior. PloS One 6(6): e20834.Google Scholar
Prat, S., Péan, S., Crépin, L. et al. 2018. The first anatomically modern humans from South-Eastern Europe. Contributions from the Buran-Kaya III Site (Crimea). Bulletins et Mémoires de la Société d’Anthropologie de Paris 30 (3–4): 169–79.Google Scholar
Preece, R. C., Gowlett, J. A. J., Parfitt, S. A., Bridgland, D. R. & Lewis, S. G.. 2006. Humans in the Hoxnian: habitat, context and fire use at Beeches Pit, West Stow, Suffolk, UK. Journal of Quaternary Science 21 (5): 485–96.Google Scholar
Premo, L. S. 2012. The shift to a predominantly logistical mobility strategy can inhibit rather than enhance forager interaction. Human Ecology 40: 647–9.Google Scholar
Premo, L. S. 2014. Cultural transmission and diversity in time-averaged assemblages. Current Anthropology 55 (1): 105–14.Google Scholar
Premo, L. S. 2016. Effective population size and the effects of demography on cultural diversity and technological complexity. American Antiquity 81 (4): 605–22.Google Scholar
Premo, L. S. & Hublin, J.-J.. 2009. Culture, population structure, and low genetic diversity in Pleistocene hominins. Proceedings of the National Academy of Sciences USA 106 (1): 33–7.Google Scholar
Proctor, C., Douka, K., Proctor, J. W. & Higham, T.. 2017. The age and context of the KC4 maxilla, Kent’s Cavern, UK. European Journal of Archaeology 20 (1): 7497.Google Scholar
Prüfer, K., Racimo, F., Patterson, N. et al. 2014. The complete genome sequence of a Neanderthal from the Altai Mountains. Nature 505: 43–9.Google Scholar
Prüfer, K., de Filippo, C., Grote, S. et al. 2017. A high-coverage Neandertal genome from Vindija Cave in Croatia. Science 358 (6363): 655–8.Google Scholar
Pryor, A. J. E., Beresford-Jones, D. G., Dudin, A. E., Ikonnikova, E. M., Hoffecker, J. F. & Gamble, C.. 2020. The chronology and function of a new circular mammoth-bone structure at Kostenki 11. Antiquity 94 (374): 323–41.Google Scholar
Qin, P. & Stoneking, M.. 2015. Denisovan ancestry in East Eurasian and Native American populations. Molecular Biology and Evolution 32: 2665–74.Google Scholar
Raghavan, M., Skoglund, P., Graf, K. E. et al. 2014. Upper Palaeolithic Siberian genome reveals dual ancestry of Native Americans. Nature 505: 8791.Google Scholar
Ramirez Rozzi, F. V. & Bermúdez de Castro, J. M. 2004. Surprisingly rapid growth in Neanderthals. Nature 428: 936–39.Google Scholar
Ramirez Rozzi, F. V., d’Errico, F, Vanhaeren, M, Grootes, P. M, Kerautret, B & Dujardin, V. 2009. Cutmarked human remains bearing Neandertal features and modern human remains associated with the Aurignacian at Les Rois. Journal of Anthropological Sciences 87: 153–85.Google Scholar
Rasmussen, S. O., Bigler, M., Blockley, S. P. et al. 2014. A stratigraphic framework for abrupt climatic changes during the Last Glacial period based on three synchronized Greenland ice-core records: refining and extending the INTIMATE event stratigraphy. Quaternary Science Reviews 106: 1428.Google Scholar
Raup, D. M. 1991. Extinction: Bad Genes or Bad Luck? New York: W.W. Norton and Company.Google Scholar
Ravon, A. L., Gaillard, C. & Monnier, J. L.. 2016. Menez-Dregan (Plouhinec, far western Europe): the lithic industry from layer 7 and its Acheulean components. Quaternary International 411: 132–43.Google Scholar
Reade, H., Tripp, J. A., Charlton, S. et al. 2020. Radiocarbon chronology and environmental context of Last Glacial Maximum human occupation in Switzerland. Scientific Reports 10: 4694.Google Scholar
Reed, D. H. 2005. Relationship between population size and fitness. Conservation Biology 19 (2): 563–8.Google Scholar
Reher, D., Key, F. M., Andrés, A. M. & Kelso, J.. 2019. Immune gene diversity in archaic and present-day humans. Genome Biology and Evolution 11 (1): 232–41.Google Scholar
Reich, D., Green, R. E., Kircher, M. et al. 2010. Genetic history of an archaic hominin group from Denisova Cave in Siberia. Nature 468: 1053–60.Google Scholar
Renard, C. 2011. Continuity or discontinuity in the Late Glacial Maximum of south-western Europe: the formation of the Solutrean in France. World Archaeology 43 (4): 726–43.Google Scholar
Revedin, A., Longo, L., Lippi, M. M. et al. 2015. New technologies for plant food processing in the Gravettian. Quaternary International 359–360: 7788.Google Scholar
Reyes-García, V., Zurro, D., Caro, J. & Madella, M. 2017. Small-scale societies and environmental transformations: coevolutionary dynamics. Ecology and Society 22 (1): 15.Google Scholar
Reynolds, N. 2020. Threading the weft, testing the warp: population concepts and the European Upper Palaeolithic chronocultural framework. In Groucutt, H. S (ed.), Culture History and Convergent Evolution: Can we Detect Populations in Prehistory? Cham, Switzerland: Springer, pp. 187212.Google Scholar
Reynolds, N. & Green, C.. 2019. Spatiotemporal modelling of radiocarbon dates using linear regression does not indicate a vector of demic dispersal associated with the earliest Gravettian assemblages in Europe. Journal of Archaeological Science: Reports 27: 101958.Google Scholar
Reynolds, N. & Riede, F.. 2019. House of cards: cultural taxonomy and the study of the European Upper Palaeolithic. Antiquity 93 (371): 1350–8.Google Scholar
Reynolds, N., Lititsyn, S. N., Sablin, M. V., Barton, N. & Higham, T. F. G.. 2015. Chronology of the European Russian Gravettian: new radiocarbon dating results and interpretation. Quartär 62: 121–32.Google Scholar
Reynolds, N., Dinnis, R., Bessudnov, A. A., Devièse, T. & Higham, T.. 2017. The Kostënki 18 child burial and the cultural and funerary landscape of Mid Upper Palaeolithic European Russia. Antiquity 91 (360): 1435–50.Google Scholar
Ribot, F., Gibert, L., Ferràndez-Cañadell, C., Olivares, E. G., Sánchez, F. & Lería, M.. 2015. Two deciduous human molars from the Early Pleistocene deposits of Barranco León (Orce, Spain). Current Anthropology 56 (1): 134–42.Google Scholar
Richards, M., Harvati, K., Grimes, V. et al. 2008. Strontium isotope evidence of Neanderthal mobility at the site of Lakonis, Greece using laser-ablation PIMMS. Journal of Archaeological Science 35: 1251–6.Google Scholar
Richards, M. P. 2009. Stable isotope evidence for European Upper Palaeolithic diets. In Hublin, J.-J & Richards, M. P. (eds.), The Evolution of Hominin Diets: Integrating Approaches to the Study of Palaeolithic Subsistence. Dordrecht: Springer, pp. 251–7.Google Scholar
Richter, D., Tostevin, G. & Škrdla, P.. 2008. Bohunician technology and thermoluminescence dating of the type locality of Brno-Bohunice (Czech Republic). Journal of Human Evolution 55: 871–85.Google Scholar
Richter, J. 2005. Hasty foragers: the Crimea Island and Europe during the last Interglacial. In Chabai, C, Richter, J & Uthmeier, T. (eds.), Kabazi II: Last Interglacial Occupation, Environment, and Subsistence: Palaeolithic Sites of Crimea, Vol. 1. Cologne: Simferopol, pp. 275–86.Google Scholar
Richter, J. 2006. Neanderthals in the landscape. In Demarsin, B. & Otte, M. (eds.), Neanderthals in Europe. Liège: ERAUL 117, pp. 5166.Google Scholar
Richter, J. 2016. Leave at the height of the party: a critical review of the Middle Palaeolithic in Western Central Europe from its beginnings to its rapid decline. Quaternary International 411: 107–28.Google Scholar
Rick, J. W. 1987. Dates as data: an examination of the Peruvian preceramic radiocarbon record. American Antiquity 52 (1): 5573.Google Scholar
Riede, F. 2009. The loss and re-introduction of bow-and-arrow technology: a case study from the Northern European Late Paleolithic. Lithic Technology 34 (1): 2745.Google Scholar
Riede, F. 2014. The resettlement of northern Europe. In Cummings, V, Jordan, P & Zvelebil, M. (eds.), Oxford Handbook of the Archaeology and Anthropology of Hunter-Gatherers. Oxford: Oxford University Press, pp. 556–81.Google Scholar
Riede., F., Johannsen, N. N., Högberg, A., Nowell, A. & Lombard, M.. 2018. The role of play objects and object play in human cognitive evolution and innovation. Evolutionary Anthropology 27 (1): 4659.Google Scholar
Riede, F., Hoggard, C. & Shennan, S.. 2019. Reconciling material cultures in archaeology with genetic data requires robust cultural evolutionary taxonomies. Palgrave Communications 5 (1): 55.Google Scholar
Riel-Salvatore, J. 2010. A niche construction perspective on the Middle-Upper Paleolithic transition in Italy. Journal of Archaeological Method and Theory 17: 323–55.Google Scholar
Riel-Salvatore, J. & Gravel-Miguel, C.. 2013. Upper Palaeolithic mortuary practices in Eurasia. In Stutz, L. N & Tarlow, S. (eds.), The Oxford Handbook of the Archaeology of Death and Burial. Oxford: Oxford University Press, pp. 303–46.Google Scholar
Riel-Salvatore, J., Gravel-Miguel, C., Maggi, R., Martino, G., Roosi, S. & Sparacello, V. S.. 2018. New insights into the Paleolithic chronology and funerary ritual of Caverna delle Arene Candide. In Borgia, V. & Cristiani, E. (eds.), Palaeolithic Italy. Advanced Studies on Early Human Adaptations in the Apennine Peninsula. Sidestone Press Academics: Leiden, pp. 335–56.Google Scholar
Rigaud, J. P. 2008. Les industries lithiques du Gravettien du nord de l’Aquitaine dans leur cadre chronologique. PALEO. Revue d’Archéologie Préhistorique (20): 381–98.Google Scholar
Rightmire, G. P., de León, M. S. P., Lordkipanidze, D., Margvelashvili, A. & Zollikofer, C. P.. 2017. Skull 5 from Dmanisi: Descriptive anatomy, comparative studies, and evolutionary significance. Journal of Human Evolution 104: 5079.Google Scholar
Riley, N. E. 2005. Demography of gender. In Poston, D. L & Micklin, M. (eds.), Handbook of Population. Dordrecht: Springer, pp. 109–41.Google Scholar
Rillardon, M. & Brugal, J.-P.. 2014. What about the Broad Spectrum Revolution? Subsistence strategy of hunter-gatherers in Southeast France between 20 and 8 ka BP. Quaternary International 337: 129–53.Google Scholar
Rink, W. J., Schwarcz, H. P., Smith, F. H. & Radovĉiĉ, J.. 1995. ESR ages for Krapina hominids. Nature 378 (6552): 24.Google Scholar
Ríos, L., Rosas, A., Estalrrich, A. et al. 2015. Possible further evidence of low genetic diversity in the El Sidrón (Asturias, Spain) Neandertal group: congenital clefts of the atlas. Plos One 10 (9): e0136550.Google Scholar
Ríos, L., Kivell, T. K., Lalueza-Fox, C. et al. 2019. Skeletal anomalies in the Neandertal family of El Sidrón (Spain) support a role of inbreeding in Neandertal extinction. Scientific Reports 9: 1697.Google Scholar
Rivero, O. & Sauvet, G.. 2014. Defining Magdalenian cultural groups in Franco-Cantabria by the formal analysis of portable artworks. Antiquity 88 (339): 6480.Google Scholar
Roberts, B. W. & Vander Linden, M. (eds.), 2011. Investigating Archaeological Cultures: Material Culture, Variability, and Transmission. New York: Springer.Google Scholar
Roberts, M. F. & Bricher, S. E.. 2018. Modeling the disappearance of the Neanderthals using principles of population dynamics and ecology. Journal of Archaeological Science 100: 1631.Google Scholar
Robinson, E., Zahid, H. J., Codding, B. F., Haas, R. & Kelly, R. L.. 2019. Spatiotemporal dynamics of prehistoric human population growth: radiocarbon ‘dates as data’ and population ecology models. Journal of Archaeological Science 101: 6371.Google Scholar
Robson, S. L. & Wood, B.. 2008. Hominin life history: reconstruction and evolution. Journal of Anatomy 212: 394425.Google Scholar
Rodríguez, J., Blain, H.-A., Mateos, A., Martín-González, J. A., Cuenca-Bescós, G. & Rodríguez-Gómez, G.. 2014. Ungulate carrying capacity in Pleistocene Mediterranean ecosystems: evidence from the Atapuerca sites. Palaeogeography, Palaeoclimatology, Palaeoecology 393: 122–34.Google Scholar
Rodríguez, J., Mateos, A., Martín-González, J. A. & Rodríguez-Gómez, G.. 2015. How rare was human presence in Europe during the Early Pleistocene? Quaternary International 389: 119–30.Google Scholar
Rodríguez, J., Mateos, A., Hertler, C. & Palombo, M. R.. 2016. Modelling human presence and environmental dynamics during the Mid-Pleistocene Revolution: new approaches and tools. Quaternary International 393: 1923.Google Scholar
Rodríguez, J., Guillermo, Z-R & Ana, M.. 2019. Does optimal foraging theory explain the behaviour of the oldest human cannibals? Journal of Human Evolution 131: 228–39.Google Scholar
Rodríguez-Gómez, G., Rodríguez, J., Martín-González, J. A., Goikoetxea, I. & Mateos, A.. 2013. Modeling trophic resource availability for the first human settlers of Europe: the case of Atapuerca TD6. Journal of Human Evolution 64: 645–57.Google Scholar
Rodríguez-Gómez, G., Mateos, A., Martín-González, J.., Blasco, R., Rosell, J. & Rodríguez, J.. 2014. Discontinuity of human presence at Atapuerca during the Early Middle Pleistocene: a matter of ecological competition? PloS One (7): e101938.Google Scholar
Rodríguez-Gómez, J. Rodríguez, J. A. Martín-González, A. Mateos, . 2017. Evaluating the impact of Homo-carnivore competition in European human settlements during the early to middle Pleistocene. Quaternary Research 88: 129–51.Google Scholar
Roebroeks, W. 2000. A marginal matter: the human occupation of northwestern Europe - 30,000 to 20,000 BP. In Roebroeks, W., Mussi, M., Svoboda, J. & Fennema, K. (eds.), Hunters of the Golden Age. The Mid Upper Palaeolithic of Eurasia 30,000–20,000 BP. Leiden: University of Leiden Press, pp. 299312.Google Scholar
Roebroeks, W. 2006. The human colonisation of Europe: where are we? Journal of Quaternary Science 21 (5): 425–35.Google Scholar
Roebroeks, W. & van Kolfschoten, T.. 1994. The earliest occupation of Europe: a short chronology. Antiquity 68: 489503.Google Scholar
Roebroeks, W., Conard, N. J. & van Kolfschoten, T.. 1992. Dense forests, cold steppes, and the Palaeolithic settlement of Northern Europe. Current Anthropology 33 (5): 551–85.Google Scholar
Roebroeks, W., Hublin, J-J & MacDonald, K.. 2011. Continuities and discontinuities in Neandertal presence: a closer look at Northwestern Europe. In Ashton, N., Lewis, S. G. & Stringer, C. (eds.), The Ancient Human Occupation of Britain. Developments in Quaternary Sciences 14, pp. 113–23.Google Scholar
Roebroeks, W., Gaudzinski-Windheuser, S., Baales, M. & Kahlke, R-D. 2018. Uneven data quality and the earliest occupation of Europe - the case of Untermassfeld (Germany). Journal of Paleolithic Archaeology 1 (1): 531.Google Scholar
Rogers, A. R., Bohlender, R. J. & Huff, C. D.. 2017. Early history of Neanderthals and Denisovans. Proceedings of the National Academy of Sciences USA 114 (37): 9859–63.Google Scholar
Rogers, A. R., Harris, N. S. & Achenbach, A. A.. 2020. Neanderthal-Denisovan ancestors interbred with a distantly related hominin. Science Advances 6 (8). http://doi.org/10.1126/sciadv.aay5483.Google Scholar
Roksandic, M. & Armstrong, S. D.. 2011. Using the life history model to set the stage(s)of growth and senescence in bioarchaeology and paleodemography. American Journal of Physical Anthropology 145: 337–47.Google Scholar
Roksandic, M., Radović, P. & Lindal, J.. 2018. Revising the hypodigm of Homo heidelbergensis: a view from the Eastern Mediterranean. Quaternary International 466 (A): 6681Google Scholar
Romanowska, I., Gamble, C., Bullock, S. & Sturt, F.. 2017. Dispersal and the Movius Line: testing the effect of dispersal on population density through simulation. Quaternary International 431: 5363.Google Scholar
Rosas, A., Martínez-Maza, C., Bastir, M. et al. 2006. Paleobiology and comparative morphology of a late Neandertal sample from El Sidrón, Spain. Proceedings of the National Academy of Sciences USA 103 (51): 19266–71.Google Scholar
Rosas, A., Ríos, L., Estalrrich, A. et al. 2017. The growth pattern of Neandertals, reconstructed from a juvenile skeleton from El Sidrón (Spain). Science 357: 1282–7.Google Scholar
Rosell, J. & Blasco, R.. 2019. The early use of fire among Neanderthals from a zooarchaeological perspective. Quaternary Science Reviews 217: 268–83.Google Scholar
Roth, E. A. 1981. Sedentism and changing fertility patterns in a Northern Athapascan isolate. Journal of Human Evolution 10: 413–25.Google Scholar
Roth, E. A. 2004. Culture, Biology, and Anthropological Demography. Cambridge: Cambridge University Press.Google Scholar
Rougier, H., Crevecoeur, I., Beauval, C. et al. 2016. Neanderthal cannibalism and Neanderthal bones used as tools in Northern Europe. Scientific Reports 6: 29005.Google Scholar
Roussel, M., Soressi, M. & Hublin, J-J. 2016. The Châtelperronian conundrum: blade and bladelet lithic technologies from Quinçay, France. Journal of Human Evolution 95: 1332.Google Scholar
Ruebens, K. 2013. Regional behaviour among late Neanderthal groups in Western Europe: a comparative assessment of late Middle Palaeolithic bifacial tool variability. Journal of Human Evolution 65: 341–62.Google Scholar
Ruebens, K. & Wragg Sykes, R. M.. 2016. Spatio-temporal variation in late Middle Palaeolithic Neanderthal behaviour: British bout coupé handaxes as a case study. Quaternary International 411: 305–36.Google Scholar
Ruebens, K., McPherron, S. J. P. & Hublin, J. J.. 2015. On the local Mousterian origin of the Châtelperonnian: integrating typo-technological, chronostratigraphic and contextual data. Journal of Human Evolution 86: 5591.Google Scholar
Sacchi, C., Schmider, B., Chantret, F. et al. 1996. Le gisement solutréen de Saint-Sulpice-de-Favières (Essonne). Bulletin de la Société Préhistorique Française 93 (4): 502–27.Google Scholar
Sala, N., Arsuaga, J. L., Martínez, I. & Gracia-Téllez, A.. 2014. Carnivore activity in the Sima de los Huesos (Atapuerca, Spain) hominin sample. Quaternary Science Reviews 97: 7183.Google Scholar
Sala, N., Arsuaga, J. L., Pantoja-Pérez, A. et al. 2015. Lethal interpersonal violence in the Middle Pleistocene. PloS One 10 (5): e0126589.Google Scholar
Sala, N., Pantoja-Pérez, A., Arsuaga, J. L., Pablos, A. & Martínez, I.. 2016. The Sima de los Huesos crania: analysis of the cranial breakage patterns. Journal of Archaeological Science 72: 2543.Google Scholar
Saladié, P., Huguet, R., Rodríguez-Hidalgo, A. et al. 2012. Intergroup cannibalism in the European Early Pleistocene: the range expansion and imbalance of power hypotheses. Journal of Human Evolution 63: 682–95.Google Scholar
Saltré, F., Brook, B. W., Rodríguez-Rey, M. et al. 2015. Uncertainties in dating constrain model choice for inferring extinction time from fossil records. Quaternary Science Reviews 112: 128–37.Google Scholar
San Juan-Foucher, C. & Foucher, P.. 2010. Marine shell beads from the Gravettian at Gargas cave (Hautes-Pyrénées, France): cultural and territorial markers. MUNIBE Suplemento – Gehigarria 31: 2835.Google Scholar
Sankararaman, S., Patterson, N., Li, H., Pääbo, S. & Reich, D.. 2012. The date of interbreeding between Neandertals and modern humans. PloS Gentics 8 (10): e1002947.Google Scholar
Sankararaman, S., Mallick, S., Dannemann, M. et al. 2014. The genomic landscape of Neanderthal ancestry in present-day humans. Nature 507: 354–7.Google Scholar
Sano, K., Arrighi, S., Stani, C. et al. 2019. The earliest evidence for mechanically delivered projectile weapons in Europe. Nature Ecology & Evolution 3 (10): 1409–14.Google Scholar
Santangelo, N., Di Donato, V., Lebreton, V., Romano, P. & Ermolli, E. R.. 2012. Palaeolandscapes of Southern Apennines during the late Early and the Middle Pleistocene. Quaternary International 267: 20–9.Google Scholar
Sattenspiel, L. & Harpending, H.. 1983. Stable populations and skeletal age. American Antiquity 48 (3): 489–98.Google Scholar
Sawyer, S., Renaud, G., Viola, B. et al. 2015. Nuclear and mitochondrial DNA sequences from two Denisovan individuals. Proceedings of the National Academy of Sciences USA 112 (51): 15696–700.Google Scholar
Scally, A. 2016. The mutation rate in human evolution and demographic inference. Current Opinion in Genetics & Development 41: 3643.Google Scholar
Scerri, E. M. L., Groucutt, H. W., Jennings, R. P. & Petraglia, M. D.. 2014. Unexpected technological heterogeneity in northern Arabia indicates complex Late Pleistocene demography at the gateway to Asia. Journal of Human Evolution 75: 125–42.Google Scholar
Scerri, E. M. L., Thomas, M. G., Manica, A. et al. 2018. Did our species evolve in subdivided populations across Africa, and why does it matter? Trends in Ecology and Evolution 33 (8): 582–94.Google Scholar
Schacht, R. 1984. The contemporaneity problem. American Antiquity 49 (4): 678–95.Google Scholar
Schiffer, M. B. 1976. Behavioural Archaeology. New York: Academic Press.Google Scholar
Schlebusch, C. M., Malmström, H., Günther, T. et al. 2017. Southern African ancient genomes estimate modern human divergence to 350,000–260,000 years ago. Science 358 (6363): 652–55.Google Scholar
Schmidt, I. & Zimmermann, A.. 2019. Population dynamics and socio-spatial organization of the Aurignacian: Scalable quantitative demographic data for western and central Europe. PloS One 14 (2): e0211562.Google Scholar
Schmidt, I., Bradtmöller, M., Kehl, M. et al. 2012. Rapid climate change and variability of settlement patterns in Iberia during the Late Pleistocene. Quaternary International 274: 179204.Google Scholar
Schmidt, I., Hilpert, J., Kretschmer, I. et al. 2021. Approaching Prehistoric Demography: proxies, scales and scope of the Cologne Protocol in European contexts. Philosophical Transactions of the Royal Society B 376: 20190714.Google Scholar
Scott, G. R. & Gibert, L.. 2009. The oldest hand-axes in Europe. Nature 461: 8285.Google Scholar
Scrimshaw, S. 1983. Infanticide as deliberate fertility regulation. In Lee, R. D. & Bulatao, R. (eds.) Determinants of Fertility in Developing Countries. Aldine: Academic Press, pp. 245–66.Google Scholar
Sear, R. 2015. Evolutionary contributions to the study of human fertility. Population Studies 69 (Supplement 1): S39S55.Google Scholar
Sear, R. & Mace, R.. 2008. Who keeps children alive? A review of the effects of kin on child survival. Evolution and Human Behavior 29: 118.Google Scholar
Seguin-Orlando, A., Korneliussen, T. S., Sikora, M. et al. 2014. Genomic structure in Europeans dating back at least 36,200 years. Science 6213 (346): 1113–18.Google Scholar
Séguy, I. & Buchet, L.. 2013. Handbook of Palaeodemography. Dordrecht: Springer.Google Scholar
Sellen, D. W. & Mace, R.. 1997. Fertility and mode of subsistence: a phylogenetic analysis. Current Anthropology 38 (5): 878–89.Google Scholar
Semal, P., Rougier, H., Crevecoeur, I. et al. 2009. New data on the Late Neandertals: direct dating of the Belgium Spy Fossils. American Journal of Physical Anthropology 138: 421–8.Google Scholar
Serangeli, J. & Bolus, M.. 2008. Out of Europe—the dispersal of a successful European hominin form. Quartär 55: 8398.Google Scholar
Serangeli, J. & Conard, N. J.. 2015. The behavioural and cultural stratigraphic contexts of the lithic assemblages from Schöningen. Journal of Human Evolution 89: 287–97.Google Scholar
Serangeli, J., Böhner, U., van Kolfschoten, T. & Conard, N. J.. 2015. Overview and new results from large-scale excavations in Schöningen. Journal of Human Evolution 89: 2745.Google Scholar
Shaffer, G., Olsen, S. M. & Bjerrum, C. J.. 2004. Ocean subsurface warming as a mechanism for coupling Dansgaard‐Oeschger climate cycles and ice‐rafting events. Geophysical Research Letters 31 (24): L24202.Google Scholar
Sharon, G. & Barsky, D.. 2016. The emergence of the Acheulian in Europe—a look from the east. Quaternary International 411: 2533.Google Scholar
Shchelinsky, V. E., Dodonov, A. E., Baigusheva, V. S. et al. 2010. Early Palaeolithic sites on the Taman Peninsula (Southern Azov Sea region, Russia): Bogatyri/Sinyaya Balka and Rodniki. Quaternary International 223–4: 2835.Google Scholar
Shchelinsky, V. E., Gurova, M., Tesakov, A. S., Titov, V. V., Frolov, P. D. & Simakova, A. N.. 2016. The Early Pleistocene site of Kermek in western Ciscaucasia (southern Russia): stratigraphy, biotic record and lithic industry (preliminary results). Quaternary International 393: 5169.Google Scholar
Shea, J. J. 2003. Neandertals, competition, and the origin of modern human behavior in the Levant. Evolutionary Anthropology 12 (4): 173–87.Google Scholar
Shea, J. J. & Sisk, M. L.. 2010. Complex projectile technology and Homo sapiens dispersal into western Eurasia. PaleoAnthropology 2010: 100–22.Google Scholar
Shennan, S. 2001. Demography and cultural innovation: a model and its implications for the emergence of modern human culture. Cambridge Archaeological Journal 11 (1): 516.Google Scholar
Shennan, S. 2002. Genes, Memes and Human History. Darwinian Archaeology and Cultural Evolution. London: Thames and Hudson.Google Scholar
Shennan, S., Downey, S. S., Timpson, A. et al. 2013. Regional population collapse followed initial agricultural booms in mid-Holocene Europe. Nature Communications 4: 2486.Google Scholar
Shipman, P. 2015. The Invaders: How Humans and Their Dogs Drove Neanderthals to Extinction. Cambridge, MA: Harvard University Press.Google Scholar
Shipton, C. & Nielsen, M.. 2015. Before cumulative culture: the evolutionary origins of overimitation and shared intentionality. Human Nature 26: 331–45.Google Scholar
Shultz, D. R., Montrey, M. & Shultz, T. R.. 2019. Comparing fitness and drift explanations of Neanderthal replacement. Proceedings of the Royal Society B 286 (1904): 20190907.Google Scholar
Sikora, M., Seguin-Orlando, A., Sousa, V. C. et al. 2017. Ancient genomes show social and reproductive behaviour of early Upper Paleolithic foragers. Science 358: 659–62.Google Scholar
Silk, J. B., Alberts, S. C. & Altmann, J.. 2003. Social bonds of female baboons enhance infant survival. Science 302 (5648): 1231–4.Google Scholar
Simonet, A. 2011. The diversity of hunting camps in the Pyrenean Gravettian. P@lethnologie, (3): 183210.Google Scholar
Simonet, A. 2017. Gravettians at Brassempouy (Landes, France), 30,000 BP: a semi-sedentary territorial organization? World Archaeology 49 (5): 648–65.Google Scholar
Simpson, G. G. 1951. The species concept. Evolution 5: 285–98.Google Scholar
Skinner, M. 1996. Developmental stress in immature hominines from Late Pleistocene Eurasia: evidence from enamel hypoplasia. Journal of Archaeological Science 23 (6): 833–52.Google Scholar
Skinner, M. 1997. Dental wear in immature Late Pleistocene European hominines. Journal of Archaeological Science 24: 677700.Google Scholar
Skoglund, P., Northoff, B. H., Shunkov, M. V. et al. 2014. Separating endogenous ancient DNA from modern day contamination in a Siberian Neandertal. Proceedings of the National Academy of Sciences USA 111 (6): 2229–34.Google Scholar
Škrdla, P. 2017. Middle to Upper Paleolithic transition in Moravia: new sites, new dates, new ideas. Quaternary International 450: 116–25.Google Scholar
Slimak, L. 2019. For a cultural anthropology of the last Neanderthals. Quaternary Science Reviews 217: 330339.Google Scholar
Slimak, L. & Giraud, Y.. 2007. Circulations sur plusieurs centaines de kilomètres durant le Paléolithique moyen. Contribution à la connnaissance des sociétés néandertaliennes. Comptes Rendus Palevol 6: 359–68.Google Scholar
Slimak, L & Nicholson, C.. 2020. Cannibals in the forest: a comment on Defleur & Desclaux (2019). Journal of Archaeological Science 117: 105034.Google Scholar
Slimak, L., Svendsen, J. L., Mangerud, J. et al. 2011. Late Mousterian persistence near the Arctic Circle. Science 332: 841–5.Google Scholar
Slon, V., Hopfe, C., Weiß, C. L. et al. 2017a. Neandertal and Denisovan DNA from Pleistocene sediments. Science 356 (6338): 605–8.Google Scholar
Slon, V., Viola, B., Renaud, G. et al. 2017b. A fourth Denisovan individual. Science Advances 3 (7): e1700186.Google Scholar
Slon, V., Mafessoni, F., Vernot, B. et al. 2018. The genome of the offspring of a Neanderthal mother and a Denisovan father. Nature 561: 113–16.Google Scholar
Smith, B. H. & Tompkins, R. L.. 1995. Toward a life history of the hominidae. Annual Review of Anthropology 24: 257–79.Google Scholar
Smith, E. A & Smith, S. A.. 1994. Inuit sex-ratio variation: population control, ethnographic error, or parental manipulation? Current Anthropology 35 (5): 595624.Google Scholar
Smith, F. H., Trinkaus, E., Pettitt, P. B., Karavanić, I. & Paunović, M.. 1999. Direct radiocarbon dates for Vindija G1 and Velika Pećina Late Pleistocene hominid remains. Proceedings of the National Academy of Sciences USA 96 (22): 12281–6.Google Scholar
Smith, F. H. 2013. The fate of the Neandertals. Journal of Anthropological Research 69: 167200.Google Scholar
Smith, F. H., Ahern, J. C. M., Janković, I. & Karavanić, I.. 2017. The assimilation model of modern human origins in light of current genetic and genomic knowledge. Quaternary International 450: 126–36.Google Scholar
Smith, P. E. L. 1966. Le Solutréen en France. Bordeaux: Delmas.Google Scholar
Smith, T. M. 2013. Teeth and human life-history evolution. Annual Review of Anthropology 42: 191208.Google Scholar
Smith, T. M., Toussaint, M., Reid, D. J ., Olejniczak, A. J. & Hublin, J.-J.. 2007. Rapid dental development in a Middle Paleolithic Belgian Neanderthal. Proceedings of the National Academy of Sciences USA 104 (51): 20220–5.Google Scholar
Smith, T. M., Harvati, K., Olejniczak, A. J., Reid, D. J., Hublin, J. J. & Panagopoulou, E.. 2009. Brief communication: dental development and enamel thickness in the Lakonis Neanderthal molar. American Journal of Physical Anthropology 138 (1): 112–18.Google Scholar
Smith, T. M., Tafforeau, P., Reid, D. J. et al. 2010. Dental evidence for ontogenetic differences between modern humans and Neanderthals. Proceedings of the National Academy of Sciences USA 107 (49): 20923–8.Google Scholar
Smith, T. M., Austin, C., Green, D. R. et al. 2018. Wintertime stress, nursing, and lead exposure in Neanderthal children. Science Advances 4 (10): eaau9483.Google Scholar
Snodgrass, J. J. & Leonard, W. R.. 2009. Neandertal energetics revisited: insights into population dynamics and life history evolution. PaleoAnthropology 2009: 220–37.Google Scholar
Soffer, O. 1985. The Upper Paleolithic of the Central Russian Plain. New York: Academic Press.Google Scholar
Soffer, O. 1987. Upper Paleolithic connubia, refugia, and the archaeological record from eastern Europe. In Soffer, O. (ed.), The Pleistocene Old World: Regional Perspectives: New York: Plenum Press, pp. 333–48.Google Scholar
Soffer, O. 1989. Storage, sedentism and the Eurasian Palaeolithic record. Antiquity 63: 719–32.Google Scholar
Soffer, O., Adovasio, J. M., Kornietz, N. J. et al. 1997. Cultural stratigraphy at Mezhirich, an Upper Palaeolithic site in Ukraine with multiple occupations. Antiquity 71: 4862.Google Scholar
Soffer, O., Adovasio, J. M. & Hyland, D. C.. 2000. The ‘Venus’ figurines: textiles, basketry, gender and status in the Upper Paleolithic. Current Anthropology 41 (4): 511–37.Google Scholar
Soficaru, A. D. & Trinkaus, E.. 2020. Perimortem versus postmortem damage: the recent case of Cioclovina 1. American Journal of Physical Anthropology 172 (1): 135–9.Google Scholar
Sommer, R. S. & Nadachowski, A.. 2006. Glacial refugia of mammals in Europe: evidence from fossil records. Mammal Review 36: 251–65.Google Scholar
Sorensen, A. C., Claud, E. & Soressi, M.. 2018. Neandertal fire-making technology inferred from microwear analysis. Scientific Reports 8: 10065.Google Scholar
Sørensen, B. 2009. Energy use in Eem Neanderthals. Journal of Archaeological Science 36: 2201–5.Google Scholar
Sørensen, B. 2011. Demography and the extinction of European Neanderthals. Journal of Anthropological Archaeology 30: 1729.Google Scholar
Sorensen, M. V. & Leonard, W. R.. 2001. Neandertal energetics and foraging efficiency. Journal of Human Evolution 40: 483–95.Google Scholar
Soressi, M. & Roussel, M.. 2014. European Middle to Upper Paleolithic industries: Châtelperronian. Encyclopaedia of Global Archaeology. New York: Springer, pp. 2679–93.Google Scholar
Soressi, M., McPherron, S. P., Lenoir, M. et al. 2013. Neandertals made the first specialized bone tools in Europe. Proceedings of the National Academy of Sciences USA 110 (35): 14186–90.Google Scholar
Spagnolo, V., Marciani, G., Aureli, D. et al. 2019. Neanderthal activity and resting areas from stratigraphic unit 13 at the Middle Palaeolithic site of Oscurusciuto (Ginosa-Taranto, Southern Italy). Quaternary Science Reviews 217: 169–93.Google Scholar
Sparacello, V. S., Rossi, S., Pettitt, P., Roberts, C., Riel-Salvatore, J. & Formicola, V.. 2018. New insights on Final Epigravettian funerary behaviour at Arene Candide cave (Western Liguria, Italy). Journal of Anthropological Sciences 96: 124.Google Scholar
Spikins, P. 2012. Goodwill hunting? Debates over the ‘meaning’ of Lower Palaeolithic handaxe form revisited. World Archaeology 44 (3): 378–92.Google Scholar
Spikins, P. A., Rutherford, H. E. & Needham, A. P.. 2010. From homininity to humanity: compassion from the earliest archaics to modern humans. Time and Mind 3 (3): 303–25.Google Scholar
Spikins, P., Hitchens, G., Needham, A. & Rutherford, H.. 2014. The cradle of thought: growth, learning, play and attachment in Neanderthal children. Oxford Journal of Archaeology 33 (2): 111–34.Google Scholar
Spikins, P., Needham, A., Tilley, L. & Hitchens, G.. 2018. Calculated or caring? Neanderthal healthcare in social context. World Archaeology 50 (3): 384403.Google Scholar
Spikins, P., Needham, A., Wright, B., Dytham, C., Gatta, M. & Hitchens, G.. 2019. Living to fight another day: the ecological and evolutionary significance of Neanderthal healthcare. Quaternary Science Reviews 217: 98–18.Google Scholar
Spinapolice, E. E. 2012. Raw material economy in Salento (Apulia, Italy): new perspectives on Neanderthal mobility patterns. Journal of Archaeological Science 39: 680–9.Google Scholar
Stade, C. M. 2020. Theory of mind as a proxy for Palaeolithic language ability: an alternative to the search for the earliest symbolic material culture. Language Dynamics and Change 10 (1): 5985.Google Scholar
Starkovich, B. M. 2017. Paleolithic subsistence strategies and changes in site use at Klissoura Cave 1 (Peloponnese, Greece). Journal of Human Evolution 111: 6384.Google Scholar
Steegman, A. T. J., Cerny, F. J. & Holliday, T. W.. 2002. Neandertal cold adaptation: physiological and energetic factors. American Journal of Physical Anthropology 14: 566–53.Google Scholar
Stewart, J. R. 2007. Neanderthal extinction as part of the faunal change in Europe during Oxygen Isotope Stage 3. Acta Zoological Cracoviensia 50A (1–2): 93124.Google Scholar
Stiner, M. C. 2013. An unshakable Middle Paleolithic? Trends versus conservatism in the predatory niche and their social ramifications. Current Anthropology 54 (8): S288304.Google Scholar
Stiner, M. C. & Kuhn, S. L.. 2006. Changes in the ‘connectedness’ and resilience of Paleolithic societies in Mediterranean ecosystems. Human Ecology 34 (5): 693712.Google Scholar
Stiner, M. C., Munro, N. D., Surovell, T. A., Tchernov, E. & Bar-Yosef, O.. 1999. Paleolithic population growth pulses evidence by small animal exploitation. Science 283: 190–4.Google Scholar
Stiner, M. C., Munro, N. D. & Surovell, T. A.. 2000. The tortoise and the hare: small game use, the broad-spectrum revolution, and Paleolithic demography. Current Anthropology 41 (1): 3973.Google Scholar
Stojanovski, D., Arzarello, M. & Nacev, T.. 2018. Middle Palaeolithic stone-tool technology from the Central Balkans: The site of Uzun Mera (eastern Republic of Macedonia). Quaternary International 476: 63–9.Google Scholar
Stoneking, M. & Krause, J.. 2011. Learning about human population history from ancient and modern genomes. Nature Reviews Genetics 12: 603–14.Google Scholar
Stout, D. & Chaminade, T.. 2012. Stone tools, language and the brain in human evolution. Philosophical Transactions of the Royal Society B 367: 7587.Google Scholar
Stout, D., Rogers, M. J., Jaeggi, A. V. & Semaw, S.. 2019. Archaeology and the origins of human cumulative culture: a case study from the earliest Oldowan at Gona, Ethiopia. Current Anthropology 60 (3): 309–40.Google Scholar
Straus, L. G. 2015. Recent developments in the study of the Upper Palaeolithic of Vasco-Cantabrian Spain. Quaternary International 364: 255–71.Google Scholar
Straus, L. G. 2018. The Upper Palaeolithic of Iberia. Trabajos de Prehistoria 75 (1): 951.Google Scholar
Straus, L. G. 2019. Just how dense on the Cantabrian Landscape were Solutrean people? Current Speculations. In Schmidt, I., Cascalheira, J., Bicho, N. & Weniger, G-C (eds.), Human Adaptations to the Last Glacial Maximum: The Solutrean and Its Neighbors. Newcastle upon Tyne: Cambridge Scholars Publishing, pp. 125.Google Scholar
Straus, L. G., Bicho, N. & Winegardner, A. C.. 2000. The Upper Palaeolithic settlement of Iberia: first-generation maps. Antiquity 74: 553–66.Google Scholar
Street, M., Jöris, O. & Turner, E.. 2012. Magdalenian settlement in the German Rhineland- an update. Quaternary International 272–3: 231–50.Google Scholar
Stringer, C. 2002. New perspectives on the Neanderthals. Evolutionary Anthropology 11 (S1): 58–9.Google Scholar
Stringer, C. 2012. The status of Homo heidelbergensis (Schoentensack 1908). Evolutionary Anthropology 21: 101–7.Google Scholar
Stringer, C. 2014. Why we are not all multiregionalists now. Trends in Ecology & Evolution 29 (5): 248–51.Google Scholar
Stringer, C. 2016. The origin and evolution of Homo sapiens. Philosophical Transactions of the Royal Society B 371: 20150237.Google Scholar
Stringer, C., Pälike, H., van Andel, T. H, Huntley, B., Valdes, P. & Allen, J. R M.. 2003. Climatic stress and the extinction of Neanderthals. In van Andel, T. H & Davies, W. (eds.), Neanderthals and Modern Humans in the European Landscape during the Last Glaciation: Archaeological Results of the Stage 3 Project. Cambridge: McDonald Institute Monographs, pp. 241–56.Google Scholar
Sullivan, A. P., de Manuel, M., Marques-Bonet, T. & Perry, G. H.. 2017. An evolutionary medicine perspective on Neandertal extinction. Journal of Human Evolution 108: 6271.Google Scholar
Surovell, T. A. & Brantingham, P. J.. 2007. A note on the use of temporal frequency distributions in studies of prehistoric demography. Journal of Archaeological Science 34: 1868–77.Google Scholar
Surovell, T. A., Finley, J. B., Smith, G. M., Brantingham, P. J. & Kelly, R.. 2009. Correcting temporal frequency distributions for taphonomic bias. Journal of Archaeological Science 36 (8): 1715–24.Google Scholar
Svensson, A., Andersen, K. K., Bigler, M. et al. 2008. A 60 000-year Greenland stratigraphic ice core chronology. Climate of the Past 4: 4757.Google Scholar
Svoboda., J.A. 2003. The Gravettian of Moravia-landscape, settlement, and dwellings. In Vasil’ev, S. A., Soffer, O. & Kozlowski, J. (eds.), Perceived Landscapes and Built Environments: The Cultural Geography of Late Palaeolithic Eurasia. Oxford: BAR International Series 1122, pp. 121–9.Google Scholar
Svoboda, J.A. 2004. Afterwords: the Pavlovian as a part of the Gravettian mosaic. In Svoboda, J. A. & Sedláčková, L. (eds.), The Gravettian along the Danube. Proceedings of the Mikulov Conference, 20–21 November 2002. Brno: Archeologický ústav AV ČR, pp. 283–97.Google Scholar
Svoboda, J. A. 2005. The Neanderthal extinction in eastern central Europe. Quaternary International 137: 6975.Google Scholar
Svoboda, J.A. 2007. The Gravettian on the Middle Danube. Paléo 19: 203220.Google Scholar
Svoboda, J. A. & Novák, A. M.. 2004. Eastern Central Europe after the Upper Pleniglacial: changing points of observation. Archäologisches Korrespondenzblatt 34 (4): 463477.Google Scholar
Svoboda, J., Ložek, V. & Vlček, E.. 1996. Hunters between East and West. The Paleolithic of Moravia. New York: Academic Press.Google Scholar
Svoboda, J., Klíma, B., Jaraošova, L. & Šrdla, P.. 2000. The Gravettian in Moravia: climate, behaviour and technological complexity. In Roebroeks, W., Mussi, M., Svoboda, J. & Fennema, K. (eds.), Hunters of the Golden Age. The Mid Upper Palaeolithic of Eurasia 30,000–20,000 BP. Leiden: University of Leiden Press, pp. 197217.Google Scholar
Svoboda, J., Péan, S. & Wojtal, P.. 2005. Mammoth bone deposits and subsistence practices during Mid-Upper Palaeolithic in Central Europe: three cases from Moravia and Poland. Quaternary International 126–8: 209–21.Google Scholar
Svoboda, J., Králík, M., Čulíková, V. et al. 2009. Pavlov VI: an Upper Palaeolithic living unit. Antiquity 83: 282–95.Google Scholar
Szymanek, M. & Julien, M-A. 2018. Early and Middle Pleistocene climate-environment conditions in Central Europe and the hominin settlement record. Quaternary Science Reviews 198: 5675.Google Scholar
Tallavaara, M., Luoto, M., Korhonen, N., Järvinen, H. & Seppä, H. 2015. Human population dynamics in Europe over the Last Glacial Maximum. Proceedings of the National Academy of Sciences USA 112 (27): 8232–7.Google Scholar
Tallavaara, M., Eronen, J. T. & Luoto, M.. 2018. Productivity, biodiversity, and pathogens influence the global hunter-gatherer population density. Proceedings of the National Academy of Sciences USA 115 (6): 1232–7.Google Scholar
Tennie, C., Call, J. & Tomasello, M.. 2009. Ratcheting up the ratchet: on the evolution of cumulative culture. Philosophical Transactions of the Royal Society B 364 (1528): 2405–15.Google Scholar
Tennie, C., Braun, D. R., Premo, L. S. & McPherron, S. P.. 2016. The island test for cumulative culture in the Paleolithic. In Haidle, M. N., Conard, N. J. & Bolus, M. (eds.), The Nature of Culture. Dordrecht: Springer, pp. 121–33.Google Scholar
Tennie, C., Premo, L. S., Braun, D. R. & McPherron, S. P.. 2017. Early stone tools and cultural transmission: resetting the null hypothesis. Current Anthropology 58 (5): 652–72.Google Scholar
Terberger, T. 2013. Le Dernier Maximum glaciaire entre le Rhin et le Danube, un réexamen critique. Mémoire LVI de la Société Préhistorique Française, 415–43.Google Scholar
Terberger, T. & Street, M.. 2002. Hiatus or continuity? New results for the question of pleniglacial settlement in Central Europe. Antiquity 76: 691–8.Google Scholar
Teschler-Nicola, M., Fernandes, D., Händel, M. et al. 2020. Ancient DNA reveals monozygotic newborn twins from the Upper Palaeolithic. Communications Biology 3: 650.Google Scholar
Teyssandier, N & Zilhão, J.. 2018. On the entity and antiquity of the Aurignacian at Willendorf (Austria): implications for modern human emergence in Europe. Journal of Paleolithic Archaeology 1: 107–38.Google Scholar
Thieme, H. 1997. Lower Palaeolithic hunting spears from Germany. Nature 385: 807–10.Google Scholar
Thompson, J. L. & Nelson, A. J.. 2011. Middle childhood and modern human origins. Human Nature 22: 249–80.Google Scholar
Tilley, L. 2015a. Theory and Practice in the Bioarchaeology of Care. Dordrecht: Springer.Google Scholar
Tilley, L. 2015b. Accommodating difference in the prehistoric past: revisiting the case of Romito 2 from a bioarchaeology of care perspective. International Journal of Paleopathology 8: 6474.Google Scholar
Tillier, A.-M. 2011. Facts and idea in Paleolithic growth studies (Paleoauxology). Evidence from Neanderthals in Europe. In Condemi, S & Weniger, G.-C. (eds.), Continuity and Discontinuity in the Peopling of Europe: One Hundred Fifty Years of Neanderthal Study. Dordrecht: Springer, pp. 139–53.Google Scholar
Timpson, A., Colledge, S., Crema, E. et al. 2014. Reconstructing regional population fluctuations in the European Neolithic using radiocarbon dates: a new case-study using an improved method. Journal of Archaeological Science 52: 549–57.Google Scholar
Toro-Moyano, I., Martínez-Navarro, B., Agustí, J. et al. 2013. The oldest human fossil in Europe, from Orce (Spain). Journal of Human Evolution 65: 19.Google Scholar
Tostevin, G. 2012. Seeing Lithics: A Middle-Range Theory for Testing for Cultural Transmission in the Pleistocene. Oxford: Oxbow.Google Scholar
Tostevin, G. & Škrdla, P.. 2006. New excavations at Bohunice and the question of the uniqueness of the type-site for the Bohunician industrial type. Anthropologie XLIV/1: 3148.Google Scholar
Tourloukis, V. 2010. The Early and Middle Pleistocene Archaeological Record of Greece: Current Status and Future Prospects. Leiden: Leiden University Press.Google Scholar
Tourloukis, V. & Harvati, K.. 2018. The Palaeolithic record of Greece: a synthesis of the evidence and a research agenda for the future. Quaternary International 466 (A): 4865.Google Scholar
Touzé, O., Flas, D. & Pesesse, D.. 2016. Technical diversity within the tanged-tool Gravettian: new results from Belgium. Quaternary International 406: 6583.Google Scholar
Trinkaus, E. 1995. Neanderthal mortality patterns. Journal of Archaeological Science 22: 121–42.Google Scholar
Trinkaus, E. 2005. The adiposity paradox in the Middle Danubian Gravettian. Anthropologie 43 (2/3): 263–72.Google Scholar
Trinkaus, E. 2011. Late Pleistocene adult mortality patterns and modern human establishment. Proceedings of the National Academy of Sciences USA 108 (4): 1267–71.Google Scholar
Trinkaus, E. 2012. Neandertals, early modern humans, and rodeo riders. Journal of Archaeological Science 39: 3691–3.Google Scholar
Trinkaus, E. 2018. An abundance of developmental anomalies and abnormalities in Pleistocene people. Proceedings of the National Academy of Sciences USA 115 (47): 11941–6.Google Scholar
Trinkaus, E. & Buzhilova, A. P.. 2012. The death and burial of Sunghir 1. International Journal of Osteoarchaeology 22: 655–66.Google Scholar
Trinkaus, E. & Svoboda, J. A.. 2006. The Paleobiology of the Pavlovian People. In Trinkaus, E & Svoboda, J. (eds.), Early Modern Human Evolution in Central Europe. The People of Dolní Věstonice and Pavlov. Oxford: Oxford University Press pp. 459–65.Google Scholar
Trinkaus, E. & Zimmerman, M. R.. 1982. Trauma among the Shanidar Neanderthals. American Journal of Physical Anthropology 57: 6176.Google Scholar
Trinkaus, E., Formicola, V., Svoboda, J., Hillson, S. W. & Holliday, T. W.. 2001. Dolní Věstonice 15: pathology and persistence in the Pavlovian. Journal of Archaeological Science 28: 1291–308.Google Scholar
Trinkaus, E., Moldovan, O., Bîlgăr, A. et al. 2003. An early modern human from the Peştera cu Oase, Romania. Proceedings of the National Academy of Sciences USA 100 (20): 11231–6.Google Scholar
Trinkaus, E., Buzhilova, A. P., Mednikova, M. B. & Dobrovolskaya, M. V.. 2014. The People of Sunghir. Burials, Bodies and Behavior in the Earlier Upper Paleolithic. Oxford: Oxford University Press.Google Scholar
Tripp, A. J. 2016. A cladistics analysis exploring regional patterning of the anthropomorphic figurines from the Gravettian. In Mendoza Straffon, L (ed.), Cultural Phlyogenetics: Concepts and Applications in Archaeology. Dordrecht: Springer, pp. 179202.Google Scholar
Turk, E. E. 2010. Hypothermia. Forensic Science, Medicine and Pathology 6: 106–15.Google Scholar
Turner, A. 1992. Large carnivores and earliest European hominids: changing determinants of resource availability during the Lower and Middle Pleistocene. Journal of Human Evolution 22 (2): 109–26.Google Scholar
Turner, C. 2000. The Eemian interglacial in the north European plain and adjacent areas. Netherlands Journal of Geosciences 79 (2/3): 217–31.Google Scholar
Turner, E., Bittmann, F., Beonigk, W. & Frechen, M.. 2000. Miesenheim I: Excavations at a Lower Palaeolithic Site in the Central Rhineland of Germany. Monographien 44. Bonn: Römich-Germanisches-Zentralmuseum.Google Scholar
Turq, A. 1999. Reflections on the Middle Palaeolithic of the Aquitaine Basin. In Roebroeks, W. & Gamble, C. (eds.), The Middle Palaeolithic Occupation of Europe. Leiden: Leiden University Press, pp. 107–20.Google Scholar
Turq, A., Faivre, J-P, Gravina, B. & Bourguignon, L.. 2017. Building models of Neanderthal territories from raw material transports in the Aquitaine Basin (southwestern France). Quaternary International 433: 88101.Google Scholar
Tzedakis, P. C., Emerson, B. C. & Hewitt, G. M.. 2013. Cryptic or mystic? Glacial tree refugia in northern Europe. Trends in Ecology and Evolution 28 (12): 696704.Google Scholar
Underdown, S. 2008. A potential role for transmissible spongiform encephalopathies in Neanderthal extinction. Medical Hypotheses 71: 47.Google Scholar
United Nations Children’s Fund and World Health Organisation. 2004. Low Birthweight. Country, Regional, and Global Estimates. New York: UNICEF.Google Scholar
United Nations Children’s Fund and World Health Organisation. 2015. Trends in Maternal Mortality: 1990 to 2015. New York: UNICEF.Google Scholar
Vaesen, K. 2012. Cumulative cultural evolution and demography. PloS One 7 (7): e40989.Google Scholar
Vaesen, K., Collard, M., Cosgrove, R. & Roebroeks, W.. 2016. Population size does not explain past changes in cultural complexity. Proceedings of the National Academy of Sciences USA 113 (16): E2241–7.Google Scholar
Vaesen, K., Scherjon, F., Hemerik, L. & Verpoorte, A.. 2019. Inbreeding, allee effects and stochasticity might be sufficient to account for Neanderthal extinction. PloS One 14 (11): e0225117.Google Scholar
Vallverdú, J., Vaquero, M., Cáceres, I. et al. 2010. Sleeping activity area within the site structure of archaic human groups: evidence from Abric Romaní Level N combustion activity areas. Current Anthropology 51 (1): 137–45.Google Scholar
van Andel, T. H. & Tzedakis, P. C.. 1996. Palaeolithic landscapes of Europe and environs, 150,000–25,000 years ago: an overview. Quaternary Science Reviews 15 (5–6): 481500.Google Scholar
van Andel, T. H., Davies, W. & Weninger, B.. 2003. The human presence in Europe during the Last Glacial Period I: human migrations and the changing climate. In van Andel, T. H. & Davies, W. (eds.), Neanderthals and Modern Humans in the European Landscape during the Last Glaciation: Archaeological Results of the Stage 3 Project. Cambridge: McDonald Institute Monographs, pp. 3156.Google Scholar
Van Arsdale, A. 2007. Preservation bias in the hominid Krapina sample? A randomization approach. Periodicum Biologorum 109 (4): 363–8.Google Scholar
Van Arsdale, P. W. 1978. Population dynamics among Asmat hunter-gatherers of New Guinea: data, methods, comparisons. Human Ecology 6 (4): 435–67.Google Scholar
van der Made, J. 2011. Biogeography and climatic change as a context to human dispersal out of Africa and within Eurasia. Quaternary Science Reviews 30 (11–12): 1353–67.Google Scholar
van der Plicht, J., Bronk Ramsey, C., Heaton, T. J., Scott, E. M. & Talamo, S.. 2020. Recent developments in calibration for archaeological and environmental samples. Radiocarbon (2020): 123.Google Scholar
van Kolfschoten, T., Buhrs, E. & Verheijen, I.. 2015. The larger mammal fauna from the Lower Paleolithic Schöningen Spear site and its contribution to hominin subsistence. Journal of Human Evolution 89: 38153.Google Scholar
van Schaik, C. P., Pradhan, G. R. & Tennie, C.. 2019. Teaching and curiosity: sequential drivers of cumulative cultural evolution in the hominin lineage. Behavioral Ecology and Sociobiology 73 (1): 2. http://doi.org/10.1007/s00265-018–2610–7.Google Scholar
Vandenberghe, J. 2002. The relation between climate and river processes, landforms and deposits during the Quaternary. Quaternary International 91 (1): 1723.Google Scholar
Vandermeersch, B. & Garralda, M. D.. 2011. Neanderthal geographical and chronological variation. In Condemi, S. & Weniger, G.-C. (eds.), Continuity and Discontinuity in the Peopling of Europe: One Hundred Fifty Years of Neanderthal Study. Dordrecht: Springer, pp. 113–25.Google Scholar
Vanhaeren, M. & d’Errico, F. 2006. Aurignacian ethno-linguistic geography of Europe revealed by personal ornaments. Journal of Archaeological Science 33 (8): 1105–28.Google Scholar
Vercellotti, G., Caramella, D., Formicola, V., Fornaciari, G. & Larsen, C. S.. 2010. Porotic hyperostosis in a Late Upper Palaeolithic skeleton (Villabruna 1, Italy). International Journal of Osteoarchaeology 20: 358–68.Google Scholar
Verna, C., Dujardin, V. & Trinkaus, E.. 2012. The Early Aurignacian human remains from La Quina-Aval (France). Journal of Human Evolution 62 (5): 605–17.Google Scholar
Vernot, B., Tucci, S., Kelso, J. et al. 2016. Excavating Neandertal and Denisovan DNA from the genomes of Melanesian individuals. Science 352 (6282): 235–39.Google Scholar
Verpoorte, A. 2003. Eastern Central Europe during the Pleniglacial. Antiquity 78: 257–66.Google Scholar
Verpoorte, A. 2006. Neanderthal energetics and spatial behaviour. Before Farming 3: 16.Google Scholar
Verpoorte, A. 2009. Limiting factors on early modern human dispersals: the human biogeography of late Pleniglacial Europe. Quaternary International 201: 7785.Google Scholar
Verpoorte, A., De Loecker, D., Niekus, M. J. L. Th & Rensink, E. 2016. The Middle Palaeolithic of the Netherlands- contexts and perspectives. Quaternary International 411: 149–62.Google Scholar
Vialet, A., Prat, S., Wils, P. & Alcicek, M. C.. 2018. The Kocabaş hominin (Denizli Basin, Turkey) at the crossroads of Eurasia: New insights from morphometric and cladistic analyses. Comptes Rendus Palevol 17 (1–2): 1732.Google Scholar
Villa, P. 1983. Terra Amata and the Middle Pleistocene Archaeological Record of Southern France, University of California Publications in Anthropology. Berkeley: University of California Press.Google Scholar
Villa, P. & Lenoir, M.. 2009. Hunting and hunting weapons of the Lower and Middle Paleolithic of Europe. In Hublin, J-J & Richards, M. P. (eds.), The Evolution of Hominin Diets. Integrating Approaches to the Study of Palaeolithic Subsistence. Springer: Dordrecht, pp. 5985.Google Scholar
Villa, P. & Roebroeks, W.. 2014. Neandertal demise: an archaeological analysis of the modern human superiority complex. PloS One 9 (4): e96424.Google Scholar
Villa, P., Pollarolo, L., Conforti, J. et al. 2018. From Neanderthals to modern humans: new data on the Uluzzian. PloS One 13 (5): e0196786.Google Scholar
Villa, P., Soriano, S., Pollarolo, L. et al. 2020. Neandertals on the beach: Use of marine resources at Grotta dei Moscerini (Latium, Italy). PloS One 15 (1): e0226690.Google Scholar
Villa, V., Pereira, A., Chaussé, C. et al. 2016. A MIS 15-MIS 12 record of environmental changes and lower Palaeolithic occupation from Valle Giumentina, central Italy. Quaternary Science Reviews 151: 160–84.Google Scholar
Villotte, S., Churchill, S. E, Dutour, O. J & Henry-Gambier, D. 2010. Subsistence activities and the sexual division of labor in the European Upper Paleolithic and Mesolithic: evidence from upper limb enthesopathies. Journal of Human Evolution 59: 3543.Google Scholar
Villotte, S., Knüsel, C. J., Mitchell, P. D. & Henry-Gambier, D.. 2011. Probable carpometacarpal and tarsal coalition from Baousso da Torre Cave (Italy): implications for burial selection during the Gravettian. Journal of Human Evolution 61: 117–20.Google Scholar
Villotte, S., Samsel, M. & Sparacello, V.. 2017. The paleobiology of two adult skeletons from Baousso da Torre (Bausu du Ture) (Liguria, Italy): implications for Gravettian lifestyle. Comptes Rendus Palevol 16: 462–73.Google Scholar
Villotte, S., Ogden, A. R. & Trinkaus, E.. 2018. Dental abnormalities and oral pathology of the Pataud 1 Upper Palaeolithic human. Bulletins et Mémoires de la Société d’Anthropologie de Paris 30: 153–61.Google Scholar
Villotte, S., Crépin, L., Rué, M. et al. 2019. Evidence for previously unknown mortuary practices in the Southwest of France (Fournol, Lot) during the Gravettian. Journal of Archaeological Science: Reports 27: 101959.Google Scholar
Viola, B. T., Gunz, P., Neubauer, S. et al. 2019. A parietal fragment from Denisova cave. Podium presentation, The 88th Annual Meeting of the American Association of Physical Anthropologists.Google Scholar
Vlček, E. 1991. Die Mammutjäger von Dolní Věstonice. Archäologie Museum 22: 136.Google Scholar
Voinchet, P., Moreno, D., Bahain, J-J et al. 2015. New chronological data (ESR and ESR/U-series) for the earliest Acheulian sites of northwestern Europe. Journal of Quaternary Science 30 (7): 610–22.Google Scholar
Volk, A. A. & Atkinson, J. A.. 2013. Infant and child death in the human environment of evolutionary adaptation. Evolution and Human Behavior 34: 182–92.Google Scholar
Wagner, G. A., Krbetschek, M., Degering, D. et al. 2010. Radiometric dating of the type-site for Homo heidelbergensis at Mauer, Germany. Proceedings of the National Academy of Sciences USA 107 (46): 19726–30.Google Scholar
Waguespack, N. M. 2002. Colonization of the Americas: disease ecology and the Paleoindian lifestyle. Human Ecology 30 (2): 227–43.Google Scholar
Wakano, J. Y., Gilpin, W., Kadowaki, S., Feldman, M. W. & Aoki, K.. 2018. Ecocultural range-expansion scenarios for the replacement or assimilation of Neanderthals by modern humans. Theoretical Population Biology 119: 314.Google Scholar
Wales, N. 2012. Modeling Neanderthal clothing using ethnographic analogues. Journal of Human Evolution 63 (6): 781–95.Google Scholar
Walker, C. S. & Churchill, S.E.. 2014. Territory size in Canis lupus: implications for Neandertal mobility. In Carlson, K. J. & Marchi, D. (eds.) Reconstructing Mobility: Environmental, Behavioral, and Morphological Determinants. Boston, MA: Springer, pp. 209–26.Google Scholar
Walker, M. 2005. Quaternary Dating Methods. Chichester: John Wiley and Sons.Google Scholar
Walker, M. J., López-Martínez, M., Carrión-Garćia, J. S. et al. 2013. Cueva Negra del Estrecho del Río Qúipar (Murcia, Spain): a late Early Pleistocene hominin site with an ‘Acheulo-Levalloiso-Mousteroid’ Palaeolithic assemblage. Quaternary International 294: 135–59.Google Scholar
Wall-Scheffler, C. M. 2012. Energetics, locomotion, and female reproduction: implications for human evolution. Annual Review of Anthropology 41: 7185.Google Scholar
Weaver, T. D. 2009. The meaning of Neandertal skeletal morphology. Proceedings of the National Academy of Sciences USA 106 (38): 16028–33.Google Scholar
Weaver, T. D. & Hublin, J. J.. 2009. Neandertal birth canal shape and the evolution of human childbirth. Proceedings of the National Academy of Sciences USA 106 (20): 8151–6.Google Scholar
Weiss, K. M. 1984. On the number of members of the genus Homo who have ever lived, and some evolutionary implications. Human Biology 56 (4): 637–49.Google Scholar
Welker, F., Hajdinjak, M., Talamo, S. et al. 2016. Palaeoproteomic evidence identifies archaic hominins associated with the Châtelperronian at the Grotte du Renne. Proceedings of the National Academy of Sciences USA 113 (40): 11162–7.Google Scholar
Wengrow, D & Graeber, D.. 2015. Farewell to the ‘childhood of man’: ritual, seasonality, and the origins of inequality. Journal of the Royal Anthropological Institute 21: 597619.Google Scholar
Weniger, G-C., de Andrés-Herrero, M, Bolin, V. et al. 2019. Late Glacial rapid climate change and human response in the Westernmost Mediterranean (Iberia and Morocco). PloS One 14 (12): e0225049.Google Scholar
Wenzel, S. 2012. The Magdalenian dwelling of Orp East (Belgium) and its spatial organisation. In Gaudzinski-Windheuser, S., Jöris, O., Sensburg, M., Street, M. & Turner, E. (eds.), Site-Internal Spatial Organization of Hunter-Gatherer Societies: Case Studies from the European Palaeolithic and Mesolithic. Verlag des Römisch-Germanischen Zentralmuseums: Mainz, pp. 141–57.Google Scholar
Westaway, K. E., Lousy, J., Awe, R. D. et al. 2017. An early modern human presence in Sumatra 73,000–63,000 years ago. Nature 548: 322–8.Google Scholar
Westaway, R. 2011. A re-evaluation of the timing of the earliest reported human occupation of Britain: the age of the sediments at Happisburgh, eastern England. Proceedings of the Geologists’ Association 122: 383–96.Google Scholar
Weyrich, L. S., Duchene, S., Soubrier, J. et al. 2017. Neanderthal behaviour, diet, and disease inferred from ancient DNA in dental calculus. Nature 544: 357–61.Google Scholar
Whallon, R. 2006. Social networks and information: non-‘utilitarian’ mobility among hunter-gatherers. Journal of Anthropological Archaeology 25 (2): 259–70.Google Scholar
White, A. 2017. A model-based analysis of the minimum size of demographically viable hunter-gatherer populations. Journal of Artificial Societies and Social Simulation 20 (4). http://doi.org/10.18564/jasss.3393.Google Scholar
White, A. A. 2014. Mortality, fertility, and the OY ratio in a model hunter-gatherer system. American Journal of Physical Anthropology 154: 222–31.Google Scholar
White, C. 1995. La Grotte du Vallonnet: evidence of early hominid activity or natural processes? Lithics 16: 70–7.Google Scholar
White, M. J. 2000. The Clactonian question: on the interpretation of core-and-flake assemblages in the British Lower Paleolithic. Journal of World Prehistory 14 (1): 1–63.Google Scholar
White, M. J. 2006. Things to do in Doggerland when you’re dead: Surviving OIS3 at the northwestern-most fringe of Middle Palaeolithic Europe. World Archaeology 38 (4): 547–75.Google Scholar
White, M. J. 2015. ‘Dancing to the rhythms of the biotidal zone’: settlement history and culture history in Middle Pleistocene Britain. In Coward, F., Hosfield, R., Pope, M. & Wenban-Smith, F. (eds.), Settlement, Society and Cognition in Human Evolution. Landscapes in Mind. Cambridge: Cambridge University Press, pp. 154–73.Google Scholar
White, M. J., Scott, B. & Ashton, N.. 2006. The Early Middle Palaeolithic in Britain: archaeology, settlement history, and human behaviour. Journal of Quaternary Science 21: 525–42.Google Scholar
White, M., Ashton, N. & Bridgland, D.. 2019. Twisted handaxes in Middle Pleistocene Britain and their implications for regional-scale cultural variation and the deep history of Acheulean hominin groups. Proceedings of the Prehistoric Society 85: 6181.Google Scholar
White, R. 2007. Systems of personal ornamentation in the Early Upper Palaeolithic: Methodological challenges and new observations. In Mellars, P., Boyle, K., Bar Yosef, O., Stringer, C. (eds.), Rethinking the Human Revolution. Cambridge: McDonald Institute Monographs, pp. 287302.Google Scholar
White, S., Gowlett, J. A. & Grove, M.. 2014. The place of the Neanderthals in hominin phylogeny. Journal of Anthropological Archaeology 35: 3250.Google Scholar
Whitehouse, R. D. 2007. Gender archaeology and the archaeology of women: do we need both? In Hamilton, S, Whitehouse, R. D. & Wright, K. I. (eds.), Archaeology and Women: Ancient and Modern Issues. Walnut Creek, CA: Left Coast Press, pp. 2740.Google Scholar
WHO (World Health Organisation). 2014. Thermal Control of the Newborn: a Practical Guide. Geneva: World Health Organisation.Google Scholar
Wiessner, P. 1974. A functional estimator of population from floor area. American Antiquity 39 (2): 343–50.Google Scholar
Wiessner, P. 2016. The rift between science and humanism: what’s data got to do with it? Current Anthropology 57 (Supplement 13): S154S166.Google Scholar
Wild, E. M., Teschler-Nicola, M., Kutschera, W., Steier, P., Trinkaus, E. & Wanek, W.. 2005. Direct dating of Early Upper Palaeolithic human remains from Mladeč. Nature 435 (7040): 332–5.Google Scholar
Wilkins, J. F. 2006. Unraveling male and female histories from human genetic data. Current Opinion in Genetics & Development 16: 611–17.Google Scholar
Willems, E. P. & van Schaik, C. P.. 2017. The social organisation of Homo ergaster: inferences from anti-predator responses in extant primates. Journal of Human Evolution 109: 1121.Google Scholar
Williams, A. N. 2012. The use of summed radiocarbon probability distributions in archaeology: a review of methods. Journal of Archaeological Science 39: 578–89.Google Scholar
Willis, K. J. & van Andel, T. H. 2004. Trees or no trees? The environments of central and eastern Europe during the Last Glaciation. Quaternary Science Reviews 23: 2369–87.Google Scholar
Wilmsen, E. N. 1982. Studies in diet, nutrition, and fertility among a group of Kalahari bushmen in Botswana. Social Science Information 21 (1): 95125.Google Scholar
Wiśniewski, A., Badura, J., Salamon, T. & Lewandowski, J.. 2014. The alleged Early Palaeolithic artefacts are in reality geofacts: a revision of the site of Kończyce Wielkie 4 in the Moravian Gate, South Poland. Journal of Archaeological Science 52: 189203.Google Scholar
Wiśniewski, A., Połtowicz-Bobak, M., Bobak, D., Jary, Z. & Moska, P.. 2017. The Epigravettian and the Magdalenian in Poland: new chronological data and an old problem. Geochronometria 44 (1): 1629.Google Scholar
Wißing, C., Rougier, H., Crevecoeur, I. et al. 2016. Isotopic evidence for dietary ecology of late Neandertals in North-Western Europe. Quaternary International 411: 327–45.Google Scholar
Wißing, C., Rougier, H., Baumann, C. et al. 2019. Stable isotopes reveal patterns of diet and mobility in the last Neanderthals and first modern humans in Europe. Scientific Reports 9: 4433.Google Scholar
Wobst, H. M. 1974. Boundary conditions for Paleolithic social systems: a simulation approach. American Antiquity 39: 147178.Google Scholar
Wobst, H. M. 1978. The archaeo-ethnology of hunter-gatherers or the tyranny of the ethnographic record in archaeology. American Antiquity 43 (2): 303–9.Google Scholar
Wolf, S. 2015. Personal ornaments as signatures of identity in the Aurignacian–the case of the Swabian Jura and western Germany. In Sanz, N. (ed.) Human Origin Sites and the World Heritage Convention in Eurasia (Vol. 2). UNESCO Publishing, pp. 92102.Google Scholar
Wolff, E. W., Chappellaz, J., Blunier, T., Rasmussen, S. O. & Svensson, A.. 2010. Millennial-scale variability during the last glacial: the ice core record. Quaternary Science Reviews 29: 2828–38.Google Scholar
Wolff, H. & Greenwood, A. D.. 2010. Did viral disease of humans wipe out the Neandertals? Medical Hypotheses 75 (1): 99105.Google Scholar
Wolpoff, M. H. & Caspari, R.. 2006. Does Krapina reflect early Neandertal paleodemography? Periodicum Biologorum 108 (4): 425–32.Google Scholar
Wood, J. W. 1990. Fertility in anthropological populations. Annual Review of Anthropology 19: 211–42.Google Scholar
Wood, J. W. 1998. A theory of preindustrial population dynamics demography, economy, and well-being in Malthusian systems. Current Anthropology 39 (1): 99135.Google Scholar
Wood, J. W., Milner, G. R., Harpending, H. C. & Weiss, K. M.. 1992. The osteological paradox: problems of inferring prehistoric health from skeletal samples. Current Anthropology 33 (4): 343–70.Google Scholar
Wood, R. E., Barroso-Ruíz, C., Caparrós, M., Jordá Pardo, J. F., Galván Santos, B. & Higham, T. F. G.. 2013. Radiocarbon dating cast doubt on the late chronology of the Middle to Upper Palaeolithic transition in southern Iberia. Proceedings of the National Academy of Sciences USA 110 (8): 2781–6.Google Scholar
Woodburn, J. 1982. Egalitarian societies. Man, 431451.Google Scholar
Wragg Sykes, R. M. 2012. Neanderthals 2.0? Evidence for expanded social networks, ethnic diversity and enculturated landscapes in the Late Middle Palaeolithic. In Ruebens, K., Romanowska, I. & Bynoe, R. (eds.), Unravelling the Palaeolithic: 10 Years of Research at the Centre for the Archaeology of Human Origins (CAHO, University of Southampton). Oxford: Archeopress, pp. 7384.Google Scholar
Wrangham, R. W. 1980. An ecological model of female-bonded primate groups. Behaviour 75 (3–4): 262300.Google Scholar
Wrangham, R. W., Wilson, M. L. & Muller, M. N.. 2006. Comparative rates of violence in chimpanzees and humans. Primates 47: 1426.Google Scholar
Wren, C. D. & Burke, A.. 2019. Habitat suitability and the genetic structure of human populations during the Last Glacial Maximum (LGM) in Western Europe. PloS One 14 (6): e0217996.Google Scholar
Wroe, S., Parr, W. C., Ledogar, J. A. et al. 2018. Computer simulations show that Neanderthal facial morphology represents adaptation to cold and high energy demands, but not heavy biting. Proceedings of the Royal Society B 285 (1876): 20180085.Google Scholar
Wygal, B. T. & Heidenreich, S. M.. 2014. Deglaciation and human colonization of northern Europe. Journal of World Prehistory 27 (2): 111–44.Google Scholar
Wynn, T., Overmann, K. A. & Coolidge, F. L.. 2016. The false dichotomy: a refutation of the Neandertal indistinguishability claim. Journal of Anthropological Sciences 94: 122.Google Scholar
Wynne-Edwards, V. C. 1962. Animal Dispersion in Relation to Social Behavior. London: Oliver & Boyd.Google Scholar
Yang, M. A. & Fu, Q.. 2018. Insights into modern human prehistory using ancient genomes. Trends in Genetics 34 (3): 184–96.Google Scholar
Yang, M. A., Gao, X., Theunert, C. et al. 2017. 40,000-year-old individual from Asia provides insight into early population structure in Eurasia. Current Biology 27: 3202–8.Google Scholar
Yellen, J. E. 1977. Archaeological Approaches to the Present: Models for Reconstructing the Past. New York: Academic Press.Google Scholar
Yokoyama, Y., Esat, T. M., Thompson, W. G. et al. 2018. Rapid glaciation and a two-step sea level plunge into the Last Glacial Maximum. Nature 559: 603–7.Google Scholar
Zahid, H. J., Robinson, E. & Kelly, R. L.. 2016. Agriculture, population growth, and statistical analysis of the radiocarbon record. Proceedings of the National Academy of Sciences USA 113 (4): 931–5.Google Scholar
Zaim, Y., Ciochon, R. L., Polanski, J. M. et al. 2011. New 1.5 million-year-old Homo erectus maxilla from Sangiran (Central Java, Indonesia). Journal of Human Evolution 61: 363–76.Google Scholar
Zhang, D., Xia, H., Chen, F. et al. 2020. Denisovan DNA in Late Pleistocene sediments from Baishiya Karst Cave on the Tibetan Plateau. Science 370: 584–7.Google Scholar
Zhou, W.-X., Sornette, D., Hill, R. A. & Dunbar, R. I. M.. 2005. Discrete hierarchical organization of social group sizes. Proceedings of the Royal Society B 272: 439–44.Google Scholar
Zhu, Z., Dennell, R., Huang, W. et al. 2018. Hominin occupation of the Chinese loess plateau since about 2.1 million years ago. Nature 559 (7715): 608–12.Google Scholar
Zhu, Z.-Y., Dennell, R., Huang, W. W. et al. 2015. New dating of the Homo erectus cranium from Lantian (Gongwangling), China. Journal of Human Evolution 78: 144–57.Google Scholar
Zihlman, A. 2013. Engendering human evolution. In Bolger, D. (ed.), A Companion to Gender Prehistory. Chichester: Wiley & Sons, pp. 2344.Google Scholar
Zilhão, J. 2005. Burial evidence for the social differentiation of age classes in the Early Upper Palaeolithic. In Vialou, D., Renault-Miskovsky, J. & Patou-Mathis, M. (eds.), Comportements des Hommes du Paléolithique Moyen et Supérieur en Europe: Territoires et Milieu. Liège: ERAUL 11, pp. 231–41.Google Scholar
Zilhão, J. 2011. Aliens from outer time? Why the ‘Human Revolution’ is wrong, and where do we go from here? In Condemi, S & Weniger, G-C (eds.), Continuity and Discontinuity in the Peopling of Europe: One Hundred Fifty Years of Neanderthal Study. Dordrecht: Springer, pp. 331–6.Google Scholar
Zilhão, J. 2013. Forty years after Roche 64: A far-west view of the Solutrean. Le Solutréen 40 ans après Smith’66. Supplément à la Revue Archéologique du Centre de la France 47, pp. 8799.Google Scholar
Zilhão, J. 2014a. The Neanderthals: evolution, palaeoecology, and extinction. In V. Cummings, P. Jordan, & M. Zvelebil (eds.), The Oxford Handbook of the Archaeology and Anthropology of Hunter-Gatherers. Oxford: Oxford University Press, pgs. 191213.Google Scholar
Zilhão, J. 2014b. The Upper Palaeolithic of Europe. In C. Renfrew & P. G. Bahn (eds.), The Cambridge World Prehistory (Vol. 3. West and Central Asia and Europe). Cambridge: Cambridge University Press, pg. 1753–85.Google Scholar
Zilhão, J. 2015. Lower and Middle Palaeolithic mortuary behaviours and the origins of ritual burial. In Renfrew, C., Boyd, M. J., & Morley, I. (eds.) Death Rituals, Social Order and the Archaeology of Immortality in the Ancient World. ‘Death Shall Have No Dominion’. Cambridge: Cambridge University Press, pp. 27-44.Google Scholar
Zilhão, J., & Trinkaus, E.. 2002. Portrait of the Artist as a Child: The Gravettian Human Skeleton from the Abrigo do Lagar Velho and its Archeological Context. Trabalhos de Arqueologia 22. Lisbon: Instituto Português de Arqueologia.Google Scholar
Zilhão, J., & Pettitt, P.. 2006. On the new dates for Gorham’s Cave and the late survival of Iberian Neanderthals. Before Farming 2006(3): 19.Google Scholar
Zilhão, J., Angelucci, D. E., Badal-García, E. et al. 2010a. Symbolic use of marine shells and mineral pigments by Iberian Neandertals. Proceedings of the National Academy of Sciences USA 107(3): 1023–8.Google Scholar
Zilhão, J., Davis, S. J., Duarte, C., Soares, A. M., Steier, P. & Wild, E.. 2010b. Pego do Diabo (Loures, Portugal): dating the emergence of anatomical modernity in westernmost Eurasia. PloS One 5(1): e8880.Google Scholar
Zilhão, J., Banks, W. E., d’Errico, F. & Gioia, P.. 2015. Analysis of site formation and assemblage integrity does not support attribution of the Uluzzian to modern humans at Grotta del Cavallo. PloS One 10 (7): e0131181.Google Scholar
Zilhão, J., Anesin, D., Aubry, T. et al. 2017. Precise dating of the Middle-to-Upper Paleolithic transition in Murica (Spain) supports late Neandertal persistence in Iberia. Heliyon 3: e00435.Google Scholar
Zilhão, J., Angelucci, D. E., Igreja, M.A. et al. 2020. Last Interglacial Iberian Neandertals as fisher-hunter-gatherers. Science 367(6485): eaaz7943.Google Scholar
Zollikofer, C. P. E., Ponce de Léon, M. S., Vandermeersch, B. & Lévêque, F.. 2002. Evidence for interpersonal violence in the St. Césaire Neanderthal. Proceedings of the National Academy of Sciences USA 99 (9): 6444–8.Google Scholar
Zubrow, E. 1989. The demographic modelling of Neanderthal extinction. In Mellars, P. & Stringer, C. (eds.), The Human Revolution. Edinburgh: Edinburgh University Press, pp. 212–31.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.

  • References
  • Jennifer C. French, University of Liverpool
  • Book: Palaeolithic Europe
  • Online publication: 27 October 2021
  • Chapter DOI: https://doi.org/10.1017/9781108590891.011
Available formats
×

Save book to Dropbox

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

  • References
  • Jennifer C. French, University of Liverpool
  • Book: Palaeolithic Europe
  • Online publication: 27 October 2021
  • Chapter DOI: https://doi.org/10.1017/9781108590891.011
Available formats
×

Save book to Google Drive

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

  • References
  • Jennifer C. French, University of Liverpool
  • Book: Palaeolithic Europe
  • Online publication: 27 October 2021
  • Chapter DOI: https://doi.org/10.1017/9781108590891.011
Available formats
×