Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-tj2md Total loading time: 0 Render date: 2024-04-18T18:46:57.869Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  03 October 2019

Alan K. Outram
Affiliation:
University of Exeter
Amy Bogaard
Affiliation:
University of Oxford
Get access
Type
Chapter
Information
Subsistence and Society in Prehistory
New Directions in Economic Archaeology
, pp. 209 - 264
Publisher: Cambridge University Press
Print publication year: 2019

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

Achilli, A., Olivieri, A., Soares, P., Lancioni, H., Kashani, B.H., Perego, U.A., et al. (2012) Mitochondrial genomes from modern horses reveal the major haplogroups that underwent domestication. Proceedings of the National Academy of Sciences 109(7), 2449–54.Google Scholar
Adams, D.C., Rohlf, F.J. and Slice, D.E. (2004) Geometric morphometrics: ten years of progress following the ‘revolution’. Italian Journal of Zoology 71(1), 516.Google Scholar
Aguilera, M., Araus, J.L., Voltas, J., Rodríguez-Ariza, M.O., Molina, F., Rovira, N., et al. (2008) Stable carbon and nitrogen isotopes and quality traits of fossil cereal grains provide clues on sustainability at the beginnings of Mediterranean agriculture. Rapid Communications in Mass Spectrometry 22(11), 1653–63.Google Scholar
Aitken, M.J. (1990) Science-Based Dating in Archaeology. London: Longman.Google Scholar
Akhinzhanov, S.M., Makarova, L.A. and Nurumov, T.N. (1992) K istorii zhivotnovodstva i ohoty v Kazakhstane (po osteologicheskomu materialu iz arheologicheskih pamiatnikov eneolita i bronzy). Almaty: Gylym.Google Scholar
Albarella, U. and Serjeantson, D. (2002) A passion for pork: meat consumption at the British late Neolithic site of Durrington Walls. In: Miracle, P. and Milner, N. (eds), Consuming Passions and Patterns of Consumption. Cambridge: McDonald Institute for Archaeological Research, pp. 3349.Google Scholar
Alexander, M.M., Gerrard, C.M., Gutiérrez, A. and Millard, A.R. (2015) Diet, society, and economy in late medieval Spain: stable isotope evidence from Muslims and Christians from Gandía, Valencia. American Journal of Physical Anthropology 156(2), 263–73.Google Scholar
Allaby, R.G. (2014) Genetics of early plant domestication: DNA and aDNA. In: Smith, C. (ed.), Encyclopedia of Global Archaeology. New York: Springer. pp. 3004–7.Google Scholar
Allard, M.W., Young, D. and Huyen, Y. (1995) Detecting dinosaur DNA. Science 268 (5214), 1192.Google Scholar
Allentoft, M.E., Sikora, M., Sjögren, K.G., Rasmussen, S., Rasmussen, M., Stenderup, J., et al. (2015) Population genomics of Bronze Age Eurasia. Nature 522(7555), 167–72.Google Scholar
Ambrose, S.H., Buikstra, J. and Krueger, H.W. (2003) Status and gender differences in diet at Mound 72, Cahokia, revealed by isotopic analysis of bone. Journal of Anthropological Archaeology 22(3), 217–26.Google Scholar
Ammerman, A.J. and Cavalli-Sforza, L.L. (1973) A population model for the diffusion of early farming in Europe. In Renfrew, C. (ed.) The Explanation of Culture Change. London: Duckworth, pp. 343–57.Google Scholar
Ananyevskaya, E., Aytqaly, A.K., Beisenov, A.Z., Dmitriev, E.A., Garbaras, A., Kukushkin, I.A., et al. (2018) Early indicators to C4 plant consumption in central Kazakhstan during the Final Bronze Age and Early Iron Age based on stable isotope analysis of human and animal bone collagen. Archaeological Research in Asia 15, 157–73.Google Scholar
Anastasiou, E. and Mitchell, P.D. (2013) Palaeopathology and genes: investigating the genetics of infectious diseases in excavated human skeletal remains and mummies from past populations. Gene 528(1), 3340.Google Scholar
Anđelinović, Š., Anterić, I., Škorić, E. and Bašić, Ž. (2015) Skeleton changes induced by horse riding on medieval skeletal remains from Croatia. The International Journal of the History of Sport 32(5), 708–21.Google Scholar
Anthony, D.W. (1996) Bridling horse power: the domestication of the horse. In: Olsen, S. (ed.), Horses through Time. Boulder, co: Roberts Rinehart, pp. 5782.Google Scholar
Anthony, D.W. (2007) The Horse, the Wheel, and Language: How Bronze-Age Riders from the Eurasian Steppes Shaped the Modern World. Princeton, nj: Princeton University Press.Google Scholar
Anthony, D.W. (2009) The Sintashta Genesis: the roles of climate change, warfare, and long-distance trade. In: Hanks, B.K., Linduff, K.M. (eds), Social Complexity in Prehistoric Eurasia. Cambridge: Cambridge University Press, pp. 4773.Google Scholar
Anthony, D.W. and Brown, D.R. (1991) The origins of horseback riding. Antiquity 65(246), 2238.Google Scholar
Anthony, D.W. and Brown, D.R. (2000) Eneolithic horse exploitation in the Eurasian steppes: diet, ritual and riding. Antiquity 74(283), 7586.Google Scholar
Anthony, D.W. and Brown, D.R. (2011) The secondary products revolution, horse-riding, and mounted warfare. Journal of World Prehistory 24(2–3), 131–60.Google Scholar
Anthony, D.W., Brown, D.R. and George, C. (2006) Early horseback riding and warfare: the importance of the magpie around the neck. In: Olsen, S.L., Grant, S., Choyke, A.M. and Bartosiewicz, L. (eds), Horses and Humans: The Evolution of Human-Equine Relationships. Oxford: Archaeopress, pp. 137–56.Google Scholar
Anthony, D.W. and Ringe, D. (2015) The Indo-European homeland from linguistic and archaeological perspectives. Annual Review of Linguistics 1(1), 199219.Google Scholar
Anthony, D., Telegin, D.Y. and Brown, D. (1991) The origin of horseback riding. Scientific American 265, 94100.Google Scholar
Araus, J.L., Febrero, A., Buxó, R., Rodríguez-Ariza, M.O., Molina, F., Camalich, M.D., et al. (1997) Identification of ancient irrigation practices based on the carbon isotope discrimination of plant seeds: a case study from the south-east Iberian Peninsula. Journal of Archaeological Science 24(8), 729–40.Google Scholar
Arbogast, R.-M. (1994) Premiers élevages néolithiques du nord-est de la France. Liège: Etudes et Recherches Archéologiques de l’Université de Liège.Google Scholar
Arnold, J.R. and Libby, W.F. (1949) Age determinations by radiocarbon content: checks with samples of known age. Science 110(2869), 678–80.Google Scholar
Asouti, E. (2013) Evolution, history and the origin of agriculture: rethinking the Neolithic (plant) economies of South-west Asia. Levant 45(2), 210–18.Google Scholar
Bailey, G., Carter, P., Gamble, C. and Higgs, H. (1983) Epirus revisited: seasonality and inter-site variation in the Upper Palaeolithic of North-West Europe. In: Bailey, G. (ed.) Hunter-Gatherer Economy in Prehistory: A European Perspective. Cambridge: Cambridge University Press, pp. 6478.Google Scholar
Baillif-Ducros, C., Truc, M.C., Paresys, C. and Villotte, S. (2012) Approche méthodologique pour distinguer un ensemble lésionnel fiable de la pratique cavalière. Exemple du squelette de la tombe 11 du site de ‘La Tuilerie’ à Saint-Dizier (Haute-Marne), VIe siècle. Bulletins et mémoires de la Société d’anthropologie de Paris 24(1–2), 2536.Google Scholar
Bakels, C.C. (1978) Four Linearbandkeramik Settlements and their Environment: a Palaeoecological Study of Sittard, Stein, Elsloo and Hienheim. Analacta Praehistorica Leidensia 11.Google Scholar
Balasse, M., Ambrose, S.H., Smith, A.B. and Price, T.D. (2002) The seasonal mobility model for prehistoric herders in the South-Western Cape of South Africa assessed by isotopic analysis of sheep tooth enamel. Journal of Archaeological Science 29, 917–32.Google Scholar
Balasse, M., Bocherens, H. and Mariotti, A. (1999) Intra-bone variability of collagen and apatite isotopic composition used as evidence of a change of diet. Journal of Archaeological Science 26, 593–8.Google Scholar
Balasse, M., Bocherens, H., Mariotti, A. and Ambrose, S.H. (2001) Detection of dietary changes by intra-tooth carbon and nitrogen isotopic analysis: an experimental study of dentine collagen of cattle (Bos taurus). Journal of Archaeological Science 28, 235–45.Google Scholar
Balasse, M. and Tresset, A. (2002) Early weaning of Neolithic domestic cattle (Bercy, France) revealed by intra-tooth variation in nitrogen isotope ratios. Journal of Archaeological Science 29(8), 853–9.Google Scholar
Ball, T., Chandler-Ezell, K., Dickau, R., Duncan, N., Hart, T.C., Iriarte, J., et al. (2016) Phytoliths as a toll for investigations of agricultural origins and dispersals around the world. Journal of Archaeological Science 68, 3245.Google Scholar
Banner, J.L. (2004) Radiogenic isotopes: systematics and applications to earth surface processes and chemical stratigraphy. Earth-Science Reviews 65(3), 141–94.Google Scholar
Barker, A., Dombrosky, J., Chaput, D., Venables, B., Wolverton, S. and Stevens, S.M. (2015) Validation of a non-targeted LC-MS approach for identifying ancient proteins: method development on bone to improve artifact residue analysis. Ethnobiology Letters 6(1), 162–74.Google Scholar
Barker, A., Dombrosky, J., Venables, B. and Wolverton, S. (2018) Taphonomy and negative results: An integrated approach to ceramic-bound protein residue analysis. Journal of Archaeological Science 94, 3243.Google Scholar
Barker, G.W.W. (1975) Prehistoric territories and economies in central Italy. In: Higgs, E.S. (ed.), Palaeoeconomy: Being the Second Volume of Papers in Economic Prehistory by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press, pp. 111–76.Google Scholar
Barker, G. (1985) Prehistoric Farming in Europe. Cambridge: Cambridge University Press.Google Scholar
Barnard, H., Ambrose, S.H., Beehr, D.E., Forster, M.D., Lanehart, R.E., Malainey, M.E., et al. (2007) Mixed results of seven methods for organic residue analysis applied to one vessel with the residue of a known foodstuff. Journal of Archaeological Science 34(1), 2837.Google Scholar
Bar-Oz, G., Weissbrod, L. and Tsahar, E. (2014) Cats in recent Chinese study on cat domestication are commensal, not domesticated. Proceedings of the National Academy of Sciences 111(10), E876.Google Scholar
Barreta, J., Gutiérrez-Gil, B., Iñiguez, V., Saavedra, V., Chiri, R., Latorre, E. and Arranz, J.J. (2013) Analysis of mitochondrial DNA in Bolivian llama, alpaca and vicuna populations: A contribution to the phylogeny of the South American camelids. Animal Genetics 44(2), 158–68.Google Scholar
Barton, H. and Fullagar, R. (2006) Microscopy. In: Torrence, R. and Barton, H. (eds) Ancient Starch Research. Walnut Creek, ca: Left Coast Press, pp. 4752.Google Scholar
Barton, H. and Torrence, R (2015) Cooking up recipes for ancient starch: assessing current methodologies and looking to the future. Journal of Archaeological Science 56, 194201.Google Scholar
Baum, T., Nendel, C., Jacomet, S., Colobran, M. and Ebersbach, R. (2016) ‘Slash and burn’ or ‘weed and manure’? A modelling approach to explore hypotheses of late Neolithic crop cultivation in pre-alpine wetland sites. Vegetation History and Archaeobotany 25, 611–27.Google Scholar
Bello, S.M. and Soligo, C. (2008) A new method for the quantitative analysis of cutmark micromorphology. Journal of Archaeological Science 35(6), 1542–52.Google Scholar
Bellone, R.R., Holl, H., Setaluri, V., Devi, S., Maddodi, N., Archer, S., et al. (2013) Evidence for a retroviral insertion in TRPM1 as the cause of congenital stationary night blindness and leopard complex spotting in the horse. PLoS One 8(10), e78280.Google Scholar
Bender, M.M., Baerreis, D.A. and Steventon, R.L. (1981) Further light on carbon isotopes and Hopewell agriculture. American Antiquity 46(2), 346–53.Google Scholar
Bendrey, R. (2007a) New methods for the identification of evidence for bitting on horse remains from archaeological sites. Journal of Archaeological Science 34(7), 1036–50.Google Scholar
Bendrey, R. (2007b) Work- and age-related changes in an Iron Age horse skeleton from Danebury hillfort, Hampshire. Archaeofauna 16, 7384.Google Scholar
Bendrey, R. (2011a) Identification of metal residues associated with bit-use on prehistoric horse teeth by scanning electron microscopy with energy dispersive X-ray microanalysis. Journal of Archaeological Science 38(11), 2989–94.Google Scholar
Bendrey, R. (2011b) Some like it hot: environmental determinism and the pastoral economies of the later prehistoric Eurasian steppe. Pastoralism 1(1), 116.Google Scholar
Bendrey, R. (2012) From wild horses to domestic horses: a European perspective. World Archaeology 44(1), 135–57.Google Scholar
Bendrey, R. (2014) Animal paleopathology. Encyclopedia of Global Archaeology. New York: Springer, pp. 258–65.Google Scholar
Bendrey, R., Hayes, T.E. and Palmer, M.R. (2009) Patterns of Iron Age horse supply: an analysis of strontium isotope ratios in teeth. Archaeometry 51, 140–50.Google Scholar
Bendrey, R., Lepetz, S., Zazzo, A., Balasse, M., Turbat, T., Giscard, P.-H., et al. (2016) Nomads, horses and mobility: an assessment of geographic origins of Iron Age horses found at Tsengel Khairkhan and Baga Turgen Gol (Mongolian Altai) based on oxygen isotope compositions of tooth enamel. In: Mashkour, M. and Beech, M. (eds), Archaeozoology of the Near East 9. Oxford: Oxbow Books, pp. 262–72.Google Scholar
Bendrey, R., Vella, D., Zazzo, A., Balasse, M. and Lepetz, S. (2014) Exponentially decreasing tooth growth rate in horse teeth: implications for isotopic analyses. Archaeometry 57(6), 1104–24.Google Scholar
Benecke, N. (1994a) Archäozoologische Studien zur Entwicklung der Haustierhaltung in Mitteleuropa und Südskandinavien von den Anfängen bis zum ausgehenden Mittelalter. Berlin: Akademie Verlag.Google Scholar
Benecke, N. (1994b) Zur Domestikation des Pferdes in Mittel- und Osteuropa: Einige neue archäozoologische Befunde. In: Hansel, B. and Zimmer, S. (eds) Die Indogermanen und das pferd. Budapest: Archaeolingua Alapítvány, pp. 123–44.Google Scholar
Benecke, N. and von den Driesch, A. (2003) Horse exploitation in the Kazakh steppes during the Eneolithic and Bronze Age. In: Levine, M., Renfrew, C. and Boyle, K. (eds) Prehistoric Steppe Adaptation and the Horse. Cambridge: McDonald Institute, pp. 6982.Google Scholar
Bentley, R.A. (2006) Strontium isotopes from the earth to the archaeological skeleton: a review. Journal of Archaeological Method and Theory 13(3), 135–87.Google Scholar
Bentley, R.A., Krause, R., Price, T.D. and Kaufmann, B. (2003) Human mobility at the early Neolithic settlement of Vaihingen, Germany: evidence from strontium isotope analysis. Archaeometry 45(3), 471–86.Google Scholar
Benz, B.F. (2001) Archaeological evidence of teosinte domestication from Guilá Naquitz, Oaxaca. Proceedings of the National Academy of Sciences 98, 2104–6.Google Scholar
Berstan, R., Stott, A.W., Minnitt, S., Ramsey, C.B., Hedges, R.E.M. and Evershed, R.P. (2008) Direct dating of pottery from its organic residues: new precision using compound-specific carbon isotopes. Antiquity 82(317), 702–13.Google Scholar
Berthon, W., Tihanyi, B., Kis, L., Révész, L., Coqueugniot, H., Dutour, O. and Pálfi, G. (2018) Horse riding and the shape of the acetabulum: insights from the bioarchaeological analysis of early Hungarian mounted archers (10th Century). International Journal of Osteoarchaeology. https://doi.org/10.1002/oa.2723Google Scholar
Bethell, P.H., Goad, L.J., Evershed, R.P. and Ottaway, J. (1994) The study of molecular markers of human activity: the use of coprostanol in the soil as an indicator of human faecal material. Journal of Archaeological Science 21(5), 619–32.Google Scholar
Bettinger, R.L. (1991) Hunter-Gatherers: Archaeological and Evolutionary Theory. New York: Plenum Press.Google Scholar
Bickle, P. and Whittle, A. (eds) (2013) The First Farmers of Central Europe. Oxford: Oxbow.Google Scholar
Bieniek, A. (2002) Archaeobotanical analysis of some early Neolithic settlements in the Kujawy region, central Poland, with potential plant gathering activities emphasised. Vegetation History and Archaeobotany 11, 3340.Google Scholar
Billamboz, A. (2006) Dendroarchäologische Untersuchungen in den neolithischen Ufersiedlungen von Hornstaad-Hörnle. In: Dieckmann, B., Harwath, A., Hoffstadt, J. and Billamboz, A. (eds), Hornstaad-Hörnle IA. Die Befunde einer jungneolithischen Pfahlbausiedlung am westlichen Bodensee. Siedlungsarchäologie im Alpenvorland IX. Stuttgart: Konrad Theiss, pp. 297414.Google Scholar
Binford, L.R. (1968) Post-Pleistocene adaptations. In: Binford, S.R. and Binford, L.R. (eds) New Perspectives in Archaeology. Chicago, il: Aldine, pp. 313–41.Google Scholar
Binford, L.R. (1978) Nunamiut Ethnoarchaeology. New York: Academic Press.Google Scholar
Binford, L.R. (1981) Bones: Ancient Men and Modern Myths. New York: Academic Press.Google Scholar
Binford, L.R. (1984) The Faunal Remains from Klasies River Mouth. New York: Academic Press.Google Scholar
Bintliff, J. (1981) Theory and reality in palaeoeconomy: some words of encouragement to the archaeologist. In: Sheridan, A. and Bailey, G. (eds), Economic Archaeology. Oxford: British Archaeological Reports, pp. 3550.Google Scholar
Bird, A. (2007) Perceptions of epigenetics. Nature 447(7143), 396–8.Google Scholar
Blackman, L. (2016) The new biologies: Epigenetics, the microbiome and immunities. Body and Society 22(4), 318.Google Scholar
Boardman, S. and Jones, G. (1990) Experiments on the effects of charring on cereal plant components. Journal of Archaeological Science 17(1), 111.Google Scholar
Boessneck, J. (1969) Osteological differences between sheep (Ovis aries Linné) and goats (Capra hircus Linné). In: Brothwell, D. and Higgs, E. (eds), Science in Archaeology. London: Thames & Hudson, pp. 331–58.Google Scholar
Bogaard, A. (2002) Questioning the relevance of shifting cultivation to Neolithic farming in the loess belt of western-central Europe: evidence from the Hambach Forest experiment. Vegetation History and Archaeobotany 11, 155–68.Google Scholar
Bogaard, A. (2004) Neolithic Farming in Central Europe. London: Routledge.Google Scholar
Bogaard, A. (2011) Plant Use and Crop Husbandry in an Early Neolithic Village: Vaihingen an der Enz, Baden-Württemberg, Frankfurter Archäologische Schriften. Bonn: Habelt-Verlag.Google Scholar
Bogaard, A. (2012) Middening and manuring in Neolithic Europe: issues of plausibility, intensity and archaeological method. In: Jones, R.L. (ed.), Manure Matters: Historical, Archaeological and Ethnographic Perspectives. Farnham: Ashgate, pp. 2539.Google Scholar
Bogaard, A., Arbogast, R.-M., Ebersbach, R., Fraser, R.A., Knipper, C., Krahn, C., et al. (2017a) The Bandkeramik settlement of Vaihingen an der Enz, Kreis Ludwigsburg (Baden-Württemberg): an integrated perspective on land use, economy and diet. Germania 94, 160.Google Scholar
Bogaard, A., Ater, M. and Hodgson, J.G. (2018) Arable weeds as a case study in plant–human relationships beyond domestication. In: Stépanoff, C. and Vigne, J.-D. (eds), Hybrid Communities: Biosocial Approaches to Domestication and Other Trans-Species Relationships. London: Routledge, pp. 97112.Google Scholar
Bogaard, A., Charles, M., Livarda, A., Ergun, M., Filipovic, D. and Jones, G. (2013) The archaeobotany of mid-later Neolithic occupation levels at Çatalhöyük. In: Hodder, I. (ed.), Humans and Landscapes of Çatalhöyük: Reports from the 2000–2008 seasons. Los Angeles: Monographs of the Cotsen Institute of Archaeology, University of California at Los Angeles, pp. 93128.Google Scholar
Bogaard, A., Fraser, R., Heaton, T.H., Wallace, M., Vaiglova, P., Charles, M., et al. (2014b) Crop manuring and intensive land management by Europe’s first farmers. Proceedings of the National Academy of Sciences 110(31), 12589–94.Google Scholar
Bogaard, A. and Halstead, P. (2015) Subsistence practices and social routine in Neolithic southern Europe. In: Fowler, C., Harding, J., Hofmann, D. (eds), The Oxford Handbook of Neolithic Europe. Oxford: Oxford University Press, pp. 385410.Google Scholar
Bogaard, A., Heaton, T.H., Poulton, P. and Merbach, I. (2007) The impact of manuring on nitrogen isotope ratios in cereals: archaeological implications for reconstruction of diet and crop management practices. Journal of Archaeological Science 34(3), 335–43.Google Scholar
Bogaard, A., Henton, E., Evans, J.A., Twiss, K.C., Charles, M.P., Vaiglova, P. and Russell, N. (2014a) Locating land use at Neolithic Çatalhöyük, Turkey: the implications of 87Sr/86Sr signatures in plants and sheep tooth sequences. Archaeometry 56(5), 860–77.Google Scholar
Bogaard, A., Hodgson, J., Nitsch, E., Jones, G., Styring, A., Diffey, C., et al. (2016) Combining functional weed ecology and crop stable isotope ratios to identify cultivation intensity: a comparison of cereal production regimes in Haute Provence, France and Asturias, Spain. Vegetation History and Archaeobotany 25, 5773.Google Scholar
Bogaard, A., Jacomet, S. and Schibler, J. (2017b) Towards an integrated bioarchaeological perspective on the central European Neolithic: understanding the pace and rhythm of social processes through comparative discussion of the western loess belt and Alpine foreland. In: Bickle, P., Cummings, V., Hofmann, D. and Pollard, J. (eds), The Neolithic of Europe. Papers in Honour of Alasdair Whittle. Oxford: Oxbow, pp. 120–42.Google Scholar
Bogaard, A. and Jones, G. (2007) Neolithic farming in Britain and central Europe: contrast or continuity? In: Whittle, A. and Cummings, V. (eds), Going Over: The Mesolithic-Neolithic Transition in North-West Europe. Proceedings of the British Academy 144. Oxford: Oxford University Press, pp. 357–75.Google Scholar
Bogaard, A., Krause, R. and Strien, H.-C. (2011) Towards a social geography of cultivation and plant use in an early farming community: Vaihingen an der Enz, south-west Germany. Antiquity 85, 395416.Google Scholar
Bogaard, A. and Outram, A.K. (2013) Palaeodiet and beyond: stable isotopes in bioarchaeology. World Archaeology 45(3), 333–7.Google Scholar
Bogaard, A., Ryan, P., Yalman, N., Asouti, E., Twiss, K.C., Mazzucato, C. and Farid, S. (2014c) Assessing outdoor activities and their social implications at Çatalhöyük. In Hodder, I. (ed.), Integrating Çatalhöyük: themes from the 2000–2008 seasons. Los Angeles: Monographs of the Cotsen Institute of Archaeology, University of California at Los Angeles, pp. 123–47.Google Scholar
Bogaard, A. and Styring, A. (2017) Plants, people and diet in the Neolithic of western Eurasia. In Lee-Thorp, J. and Katzenberg, A. (eds), Oxford Handbook of Ancient Diet. Oxford: Oxford University Press. DOI:10.1093/oxfordhb/9780199694013.013.45.Google Scholar
Bogaard, A., Styring, A., Ater, M., et al. (2018) From traditional farming in Morocco to early urban agroecology in northern Mesopotamia: combining present-day arable weed surveys and crop isotope analysis to reconstruct past agrosystems in (semi-) arid regions. Environmental Archaeology 23, 30322. DOI: 10.1080/14614103.2016.1261217.Google Scholar
Bogucki, P.I. (1984) Ceramic sieves of the Linear Pottery Culture and their economic implications. Oxford Journal of Archaeology 3(1), 1530.Google Scholar
Boivin, N., Fuller, D.Q. and Crowther, A. (2012). Old World globalization and the Columbian exchange: Comparison and contrast. World Archaeology 44(3), 452–69.Google Scholar
Bökönyi, S. (1974) History of Domestic Animals in Central and Eastern Europe. Budapest: Akadémiai Kiadó.Google Scholar
Bollongino, R., Edwards, C.J., Alt, K.W., Burger, J. and Bradley, D.G. (2006) Early history of European domestic cattle as revealed by ancient DNA. Biology Letters 2, 155–9.Google Scholar
Bollongino, R., Elsner, J., Vigne, J.-D. and Burger, J. (2008) Y-SNPs do not indicate hybridisation between European aurochs and domestic cattle. PloS ONE 3, e3418.Google Scholar
Bolton, L.S. (2007) An investigation into why a selection of Bronze Age pots from Kazakhstan have been repaired with bronze staples. Unpublished undergraduate dissertation, University of Exeter.Google Scholar
Boserup, E. (1965) The Conditions of Agricultural Growth. Woking and London: Unwin Brothers.Google Scholar
Boserup, E. (1981) Population and Technology. Oxford: Blackwell.Google Scholar
Bouckaert, R., Lemey, P., Dunn, M., Greenhill, S.J., Alekseyenko, A.V., et al. (2012) Mapping the origins and expansion of the Indo-European language family. Science 337, 957–60.Google Scholar
Bowen, G.J. (2010) Isoscapes: spatial pattern in isotopic biogeochemistry. Annual Review of Earth and Planetary Sciences 38, 161–87.Google Scholar
Bowen, G.J. and Wilkinson, B. (2002) Spatial distribution of δ18O in meteoric precipitation. Geology 30, 315–18.Google Scholar
Bradley, D.G. (2006) Documenting domestication: reading animal genetic texts. In: Zeder, M.A., Bradley, D.G., Emschwiller, E. and Smith, B.D. (eds), Documenting Domestication: New Genetic and Archaeological Paradigms. Berkeley: University of California Press, pp. 273–8.Google Scholar
Bradley, D.G., MacHugh, D.E., Cunningham, P. and Loftus, R.T. (1996) Mitochondrial diversity and the origins of African and European cattle. Proceedings of the National Academy of Sciences 93(10), 5131–5.Google Scholar
Bradley, R., Haselgrove, C., Vander Linden, M. and Webley, L. (2016) The Later Prehistory of Western Europe: The Evidence of Development-Led Fieldwork. Oxford: Oxford University Press.Google Scholar
Bramanti, B., Thomas, M.G., Haak, W., Unterländer, M., Jores, P., Tambets, K., et al. (2009) Genetic discontinuity between local hunter-gatherers and central Europe’s first farmers. Science 326(5949), 137–40.Google Scholar
Brill, R.H. and Wampler, J.M. (1965) Isotope ratios in archaeological objects of lead. In: Application of science in examination of works of art. Proceedings of the seminar: September 7–16, 1965. Boston, ma: Museum of Fine Arts, pp. 155–66.Google Scholar
Brill, R.H. and Wampler, J.M. (1967) Isotope studies of ancient lead. American Journal of Archaeology 71(1), 6377.Google Scholar
Britton, K., Grimes, V., Niven, L., Steele, T.E., McPherron, S., Soressi, M., et al. (2011) Strontium isotope evidence for migration in late Pleistocene Rangifer: implications for Neanderthal hunting strategies at the Middle Palaeolithic site of Jonzac, France. Journal of Human Evolution 61(2), 176–85.Google Scholar
Brochier, J.É. (2013) The use and abuse of culling profiles in recent zooarchaeological studies: some methodological comments on ‘frequency correction’ and its consequences. Journal of Archaeological Science 40(2), 1416–20.Google Scholar
Brombacher, C. (1997) Archaeobotanical investigations of Late Neolithic lakeshore settlements (Lake Biel, Switzerland). Vegetation History and Archaeobotany 6, 167–86.Google Scholar
Bronson, B. 1972. Farm labor and the evolution of food production. In: Spooner, B. (ed.), Population Growth: Anthropological Implications. Cambridge, ma: MIT Press, pp. 190218.Google Scholar
Brown, D. and Anthony, D. (1998) Bit wear, horseback riding and the Botai site in Kazakhstan. Journal of Archaeological Science 25(4), 331–47.Google Scholar
Brown, T.A. (2001) Ancient DNA. In: Brothwell, D.R. and Pollard, A.M. (eds), Handbook of Archaeological Sciences. Chichester: Wiley, pp. 301–12.Google Scholar
Brown, T. and Brown, K. (2011) Biomolecular Archaeology: An Introduction. Oxford: Wiley-Blackwell.Google Scholar
Buckland, P.C., Amorosi, T., Barlow, L.K., Dugmore, A.J., Mayewski, P.A., McGovern, T.H., et al. (1996) Bioarchaeological and climatological evidence for the fate of Norse farmers in medieval Greenland. Antiquity 70(267), 8896.Google Scholar
Buckley, M., Fraser, S., Herman, J., Melton, N.D., Mulville, J. and Pálsdóttir, A.H. (2014) Species identification of archaeological marine mammals using collagen fingerprinting. Journal of Archaeological Science 41, 631–41.Google Scholar
Buckley, M., Kansa, S.W., Howard, S., Campbell, S., Thomas-Oates, J. and Collins, M. (2010) Distinguishing between archaeological sheep and goat bones using a single collagen peptide. Journal of Archaeological Science 37(1), 1320.Google Scholar
Budd, P., Montgomery, J., Evans, J. and Barreiro, B. (2000) Human tooth enamel as a record of the comparative lead exposure of prehistoric and modern people. Science of the Total Environment 263(1–3), 110.Google Scholar
Bull, I.D., Simpson, I.A., van Bergen, P.F. and Evershed, R.P. (1999) Muck ‘n’ molecules: organic geochemical methods for detecting ancient manuring. Antiquity 73(279), 8696.Google Scholar
Bunning, S.L., Jones, G. and Brown, T.A. (2012) Next generation sequencing of DNA in 3300-year-old charred cereal grains. Journal of Archaeological Science 39(8), 2780–4.Google Scholar
Burger, O., Hamilton, M.J. and Walker, R. (2005) The prey as patch model: optimal handling of resources with diminishing returns. Journal of Archaeological Science 32(8), 1147–58.Google Scholar
Burger, J., Kirchner, M., Bramanti, B., Haak, W. and Thomas, M.G. (2007) Absence of the lactase-persistence-associated allele in early Neolithic Europeans. Proceedings of the National Academy of Sciences 104, 3736–41.Google Scholar
Burton, J.H. and Hahn, R. (2016) Assessing the ‘local’ 87Sr/86Sr ratio for humans. In: Grupe, G. and McGlynn, G.V. (eds) Isotopic Landscapes in Bioarchaeology. Heidelberg: Springer, pp. 113–21.Google Scholar
Buttler, W. and Haberey, W. (1936) Die Bandkeramische Ansiedlung bei Köln-Lindenthal. Berlin: de Gruyter.Google Scholar
Buxó, R. (2007) Early agriculture in central and southern Spain. In: Colledge, S. and Conolly, J. (eds), The Origins and Spread of Domestic Plants in Southwest Asia and Europe. Walnut Creek, ca: Left Coast Press, pp. 155–71.Google Scholar
Campos, P.F., Willerslev, E., Sher, A., Orlando, L., Axelsson, E., Tikhonov, A., et al. (2010) Ancient DNA analyses exclude humans as the driving force behind late Pleistocene musk ox (Ovibos moschatus) population dynamics. Proceedings of the National Academy of Sciences 107(12), 5675–80.Google Scholar
Cannon, D.Y. (1987) Marine Fish Osteology: A Manual for Archaeologists. Burnaby: Simon Fraser University.Google Scholar
Cardini, A., Seetah, K. and Barker, G. (2015) How many specimens do I need? Sampling error in geometric morphometrics: Testing the sensitivity of means and variances in simple randomized selection experiments. Zoomorphology 134(2), 149–63.Google Scholar
Carniero, R.L. (1960) Slash-and-burn agriculture: a closer look at its implications for settlement patterns. In: Wallace, A.F.C. (ed.), Men and Cultures. Philadelphia: University of Pennsylvania Press, pp. 229–34.Google Scholar
Carter, R.J. (1997) Age estimation of the roe deer (Capreolus capreolus) mandibles from the Mesolithic site of Star Carr, Yorkshire, based on radiographs of mandibular tooth development. Journal of Zoology 241(3), 495502.Google Scholar
Carter, R.J. (1998) Reassessment of seasonality at the early Mesolithic site of Star Carr, Yorkshire based on radiographs of mandibular tooth development in red deer (Cervus elaphus). Journal of Archaeological Science 25(9), 851–6.Google Scholar
Casanova, E., Knowles, T.D., Williams, C., Crump, M.P. and Evershed, R.P. (2018) Practical considerations in high-precision compound-specific radiocarbon analyses: eliminating the effects of solvent and sample cross-contamination on accuracy and precision. Analytical Chemistry 90(18), 11025–32.Google Scholar
Cashdan, E. (1983a) Territoriality among human foragers: ecological models and an application to four Bushman groups. Current Anthropology 24(1), 4755.Google Scholar
Cashdan, E. (1983b) Reply. Current Anthropology 24(1), 62–6.Google Scholar
Castillo, C. (2011) Rice in Thailand: the archaeobotanical contribution. Rice 4, 114–20.Google Scholar
Caulfield, S. (1978) Star Carr – an alternative view. Irish Archaeological Research Forum 5, 1522.Google Scholar
Chang, C. (2006) The grass is greener on the other side: a study of pastoral mobility on the Eurasian steppe of southeastern Kazakhstan. In: Sellet, F., Greaves, R. and Yu, P.-L. (eds), Archaeology and Ethnoarchaeology of Mobility. Gainesville: University Press of Florida, pp. 184200.Google Scholar
Chaplin, R.E. (1971) The Study of Animal Bones from Archaeological Sites. London: Seminar Press.Google Scholar
Charles, M., Bogaard, A., Jones, G., Hodgson, J. and Halstead, P. (2002) Ecological investigation of intensive cereal cultivation in the mountains of Asturias, NW Spain. Vegetation History and Archaeobotany 11, 133–42.Google Scholar
Charles, M., Forster, E., Wallace, M. and Jones, G. (2015) ‘Nor ever lightning char thy grain’: establishing archaeologically relevant charring conditions and their effect on glume wheat grain morphology. Science and Technology of Archaeological Research 1, DOI: 10.1179/2054892315Y.0000000008.Google Scholar
Charnov, E.L. (1976) Optimal foraging, the marginal value theorem. Theoretical Population Biology 9(2), 129–36.Google Scholar
Chen, Y.S., Torroni, A., Excoffier, L., Santachiara-Benerecetti, A.S. and Wallace, D.C. (1995) Analysis of mtDNA variation in African populations reveals the most ancient of all human continent-specific haplogroups. American Journal of Human Genetics 57(1), 133–49.Google Scholar
Childe, V.G. (1929) The Danube in Prehistory. Oxford: Clarendon Press.Google Scholar
Childe, V.G. (1954) What Happened in History. Harmsondsworth: Penguin.Google Scholar
Childe, V.G. (1957) The Dawn of European Civilization. London: Routledge.Google Scholar
Chisholm, M. (1962) Rural Settlement and Land Use. London: Hutchinson.Google Scholar
Clark, G. (1939) Archaeology and Society: Reconstructing the Prehistoric Past. London: Methuen.Google Scholar
Clark, G. (1957) Archaeology and Society: Reconstructing the Prehistoric Past, 3rd edn. London: Methuen.Google Scholar
Clark, J.D.G. (1942) Bees in antiquity. Antiquity 16, 208–15.Google Scholar
Clark, J.D.G. (1947a) Forest clearance and prehistoric farming. Economic History Review, 17, 4551.Google Scholar
Clark, J.D.G. (1947b) Whales as an economic factor in prehistoric Europe. Antiquity 21, 84104.Google Scholar
Clark, J.D.G. (1952) Prehistoric Europe: The Economic Basis. London: Methuen.Google Scholar
Clark, J.D.G. (1953) The economic approach to prehistory: Albert Reckitt Archaeological lecture 1953. Proceedings of the British Academy 39, 215–38.Google Scholar
Clark, J.D.G. (1954) Excavations at Star Carr: An Early Mesolithic Site at Seamer Near Scarborough, Yorkshire. Cambridge: Cambridge University Press.Google Scholar
Clark, J.D.G. (1961) World Prehistory: An Outline. Cambridge: Cambridge University Press.Google Scholar
Clark, J.D.G. (1972a) Foreword. In: Higgs, E.S. (ed.), Papers in Economic Prehistory: Studies by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press, pp. viix.Google Scholar
Clark, J.D.G. (1972b) Star Carr: A Case Study in Bioarchaeology. Reading, ma: Addison-Wesley.Google Scholar
Clark, J.D.G. (1989) Prehistory and Cambridge and Beyond. Cambridge: Cambridge University Press.Google Scholar
Clarke, D. (1976) Mesolithic Europe: the economic basis. In: de G. Sieveking, D., Longworth, I.H. and Wilson, K.E. (eds), Problems in Economic and Social Archaeology. London: Duckworth, pp. 449–81.Google Scholar
Claβen, E. (2011) Siedlungen der Bandkeramik bei Königshoven. Rheinische Ausgrabungen 64.Google Scholar
Clayton, F., Sealy, J. and Pfeiffer, S. (2006) Weaning age among foragers at Matjes River Rock Shelter, South Africa, from stable nitrogen and carbon isotope analyses. American Journal of Physical Anthropology 129(2), 311–17.Google Scholar
Codding, B.F. and Bird, D.W. (2015) Behavioral ecology and the future of archaeological science. Journal of Archaeological Science 56, 920.Google Scholar
Cohen, A. and Serjeantson, D. (1986) A Manual for the Identification of Bird Bones from Archaeological Sites. London: Archetype.Google Scholar
Coles, J. (1973) Archaeology by Experiment. London: Hutchinson.Google Scholar
Coles, J. (1979) Experimental Archaeology. London: Academic Press.Google Scholar
Collins, M., Buckley, M., Grundy, H.H., Thomas-Oates, J., Wilson, J. and van Doorn, N. (2010) ZooMS: the collagen barcode and fingerprints. Spectroscopy Europe 22(2), 6.Google Scholar
Collins, M.J. and Copeland, L. (2011) Ancient starch: Cooked or just old? Proceedings of the National Academy of Sciences 108, E145.Google Scholar
Condamin, J., Formenti, F., Metais, M.O., Michel, M. and Blond, P. (1976) The application of gas chromatography to the tracing of oil in ancient amphorae. Archaeometry 18(2), 195201.Google Scholar
Conklin, H.C. (1959) Population-land balance under systems of tropical forest agriculture. Proceedings Pacific Science Congress 7, 63.Google Scholar
Cook, W.R. (2011) Damage by the bit to the equine interdental space and second lower premolar. Equine Veterinary Education 23(7), 355–60.Google Scholar
Cooper, A. and Poinar, H.N. (2000) Ancient DNA: do it right or not at all. Science 5482(1139), 416.Google Scholar
Copley, M.S., Berstan, R., Dudd, S.N., Docherty, G., Mukherjee, A. J., Straker, V., Payne, S. and Evershed, R.P. (2003) Direct chemical evidence for widespread dairying in prehistoric Britain. PNAS 100, 1524–9.Google Scholar
Copley, M.S., Berstan, R., Mukherjee, A.J., Dudd, S.N., Straker, V., Payne, S. and Evershed, R.P. (2005a) Dairying in antiquity. III. Evidence from absorbed lipid residues dating to the British Neolithic. Journal of Archaeological Science 32, 523–46.Google Scholar
Copley, M.S., Berstan, R., Straker, V., Payne, S. and Evershed, R.P. (2005b) Dairying in antiquity. II. Evidence from absorbed lipid residues dating to the British Bronze Age. Journal of Archaeological Science 32, 505–21.Google Scholar
Copley, M.S., Berstan, R., Dudd, S. N., Straker, V., Payne, S. and Evershed, R.P. (2005c) Dairying in antiquity. I. Evidence from absorbed lipid residues dating to the British Iron Age. Journal of Archaeological Science 32, 485503.Google Scholar
Copley, M.S., Evershed, R.P., Rose, P.J., Clapham, A., Edwards, D.N. and Horton, M.C. (2001) Processing palm fruits in the Nile Valley – biomolecular evidence from Qasr Ibrim. Antiquity 75, 538–42.Google Scholar
Corr, L.T., Richards, M.P., Jim, S., Ambrose, S.H., Mackie, A. and Evershed, R.P. (2009) Probing dietary change of Kwädąy Dän Ts’ìnchı̨, an ancient glacier body from British Columbia: II. Deconvoluting whole skin and bone collagen δ13C values via carbon isotope analysis of individual amino acids. Journal of Archaeological Science 36, 1218.Google Scholar
Craig, H. (1953) The geochemistry of the stable carbon isotopes. Geochimica et Cosmochimica Acta 3(2–3), 5392.Google Scholar
Craig, O.E., Chapman, J., Figler, A., Patay, P., Taylor, G. and Collins, M.J. (2003) ‘Milk jugs’ and other myths of the copper age of central Europe. European Journal of Archaeology 6(3), 251–65.Google Scholar
Craig, O.E., Forster, M., Andersen, S.H., Koch, E., Crombé, P., Milner, N.J., et al. (2007) Molecular and isotopic demonstration of the processing of aquatic products in northern European prehistoric pottery. Archaeometry 49(1), 135–52.Google Scholar
Craig, O.E., Love, G.D., Isaksson, S., Taylor, G. and Snape, C.E. (2004) Stable carbon isotopic characterisation of free and bound lipid constituents of archaeological ceramic vessels released by solvent extraction, alkaline hydrolysis and catalytic hydropyrolysis. Journal of Analytical and Applied Pyrolysis 71(2), 613–34.Google Scholar
Craig, O., Mulville, J., Pearson, M.P., Sokol, R., Gelsthorpe, K., Stacey, R. and Collins, M. (2000) Archaeology: detecting milk proteins in ancient pots. Nature 408(6810), 312.Google Scholar
Cramp, L.J.E. and Evershed, R.P. (2014) Reconstructing aquatic resource exploitation in human prehistory using lipid biomarkers and stable isotopes. In: Holland, H.D. and Turekian, K.K. (eds), Treatise on Geochemistry: Archaeology and Anthropology. Oxford and Amsterdam: Elsevier, pp. 319–39Google Scholar
Cramp, L.J., Jones, J., Sheridan, A., Smyth, J., Whelton, H., Mulville, J., et al. (2014) Immediate replacement of fishing with dairying by the earliest farmers of the northeast Atlantic archipelagos. Proceedings of the Royal Society B: Biological Sciences 281(1780), 20132372.Google Scholar
Crowther, A. (2012) The differential survival of native starch during cooking and implications for archaeological analyses: a review. Archaeological and Anthropological Sciences 4, 221–35.Google Scholar
Crowther, A., Haslam, M., Oakden, N., Walde, D. and Mercader, J. (2014) Documenting contamination in ancient starch laboratories. Journal of Archaeological Science 49, 90104.Google Scholar
Cucchi, T., Mohaseb, A., Peigné, S., Debue, K., Orlando, L. and Mashkour, M. (2017) Detecting taxonomic and phylogenetic signals in equid cheek teeth: towards new palaeontological and archaeological proxies. Royal Society Open Science 4(4), 160997.Google Scholar
Cunliffe, B. (2015) By Steppe, Desert, and Ocean: The Birth of Eurasia. Oxford: Oxford University Press.Google Scholar
Currat, M. and Excoffier, L. (2004) Modern humans did not admix with Neanderthals during their range expansion into Europe. PLoS Biology 2(12), e421.Google Scholar
Dahm, R. (2008) Discovering DNA: Friedrich Miescher and the early years of nucleic acid research. Human Genetics 122(6), 565–81.Google Scholar
Damgaard, P.d.B., Marchi, N., Rasmussen, S., Peyrot, M., Renaud, G., Korneliussen, T., et al. (2018b) 137 ancient human genomes from across the Eurasian steppes. Nature 557(7705), 369–74.Google Scholar
Damgaard, P.d.B., Martiniano, R., Kamm, J., Moreno-Mayar, J.V., Kroonen, G., Peyrot, M., et al. (2018a) The first horse herders and the impact of Early Bronze Age steppe expansions into Asia. Science, DOI: 10.1126/science.aar7711.Google Scholar
Dansgaard, W. (1964) Stable isotopes in precipitation. Tellus 16(4), 436–68.Google Scholar
Davidson, I. (1981) Can we study prehistoric economy for fisher-gatherer-hunters? A historical approach to Cambridge ‘palaeoeconomy’. In: Sheridan, A. and Bailey, G. (eds), Economic Archaeology. Oxford: British Archaeological Reports, pp. 1733.Google Scholar
Davis, S.J. (1987) The Archaeology of Animals. London: Batsford.Google Scholar
Delhon, C., Martin, L., Argant, J. and Thiébault, S. (2008) Shepherds and plants in the Alps: multi-proxy archaeobotanical analysis of Neolithic dung from ‘La Grande Rivoire’ (Isère, France). Journal of Archaeological Science 35, 2937–52.Google Scholar
Denham, T.P., Haberle, S.G., Lentfer, C., Fullagar, R., Field, J., Therin, M., et al. (2003) Origins of agriculture at Kuk Swamp in the Highlands of New Guinea. Science 301, 189–93.Google Scholar
Denhardt, D. (2017) Effect of stress on human biology: epigenetics, adaptation, inheritance and social significance. Journal of Cellular Physiology. DOI: 10.1002/jcp.25837.Google Scholar
DeNiro, M.J. and Epstein, S. (1976) You are what you eat (plus a few ‰): the carbon isotope cycle in food chains. Geological Society of America 6, 834.Google Scholar
DeNiro, M.J. and Epstein, S. (1978) Influence of diet on the distribution of carbon isotopes in animals. Geochimica et Cosmochimica Acta 42(5), 495506.Google Scholar
DeNiro, M.J. and Epstein, S. (1981) Influence of diet on the distribution of nitrogen isotopes in animals. Geochimica et Cosmochimica Acta 45(3), 341–51.Google Scholar
Dennell, R.W. (1972) The interpretation of plant remains: Bulgaria. In: Higgs, E.S. (ed.), Papers in Economic Prehistory: Studies by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press, pp. 149–59.Google Scholar
Dennell, R.W. (1976) The economic importance of plant resources represented on archaeological sites. Journal of Archaeological Science 3(3), 229–47.Google Scholar
Dennell, R.W. (1983) European Economic Prehistory: A New Approach. London: Academic Press.Google Scholar
Dennell, R.W. (1985) The hunter/gatherer/agricultural frontier in prehistoric temperate Europe. In: Green, S. and Perlman, S.M. (eds), The Archaeology of Frontiers and Boundaries. New York: Academic Press, pp. 113–40.Google Scholar
Dennell, R.W. and Webley, D. (1975) Prehistoric settlement and land use in southern Bulgaria. In: Higgs, E.S. (ed.), Palaeoeconomy: Being the Second Volume of Papers in Economic Prehistory by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press, pp. 97110.Google Scholar
Deschler-Erb, S. and Marti-Grädel, E. (2004) Viehhaltung und Jagd. Ergebnisse der Untersuchung der handaufgelesenen Tierknochen. In: Jacomet, S., Leuzinger, U. and Schibler, J. (eds), Die jungsteinzeitliche Seeufersiedlung Arbon Bleiche 3: Umwelt und Wirtschaft. Frauenfeld: Amt für Archäologie des Kantons Thurgau, pp. 158252.Google Scholar
Deschler-Erb, S., Marti-Grädel, E. and Schibler, J. (2002) Bukranien in der jungsteinzeitlichen Siedlung Arbon-Bleiche 3: Status, Kult oder Zauber? Archäologie der Schweiz 25: 2533.Google Scholar
Dickau, R., Bruno, M.C., Iriarte, J., Prümers, H., Betancourt, C.J., Holst, I. and Mayle, F.E. (2012) Diversity of cultivars and other plant resources used at habitation sites in the Llanos de Mojos, Beni, Bolivia: evidence from macrobotanical remains, starch and phytoliths. Journal of Archaeological Science 39, 357–70.Google Scholar
Dieckmann, B., Maier, U. and Vogt, R. (2001) Hornstaad – Zur inneren Dynamik einer jungneolithischen Dorfanlage am westlichen Bodensee. In: Lippert, A., Schultz, M., Shennan, S. and Teschler-Nicola, M. (eds) Mensch und Umwelt während des Neolithikums und der Frühbronzezeit in Mitteleuropa. Rahden: Verlag Marie-Leidorf GmbH, pp. 2951.Google Scholar
Doebley, J. (2004) The genetics of maize evolution. Annual Review of Genetics 38, 3759.Google Scholar
Döhle, H.-J. (1997). Zum Stand der Untersuchungen an neolithischen Tierknochen aus Mitteldeutschland. Jahresschrift für mitteldeutsche Vorgeschichte 79: 111–47.Google Scholar
Dole, M., Lane, G.A., Rudd, D.P. and Zaukelies, D.A. (1954) Isotopic composition of atmospheric oxygen and nitrogen. Geochimica et Cosmochimica Acta 6(2), 6578.Google Scholar
Dong, Y., Li, C., Luan, F., Li, Z., Li, H., Cui, Y., et al. (2015) Low mitochondrial DNA diversity in an ancient population from China: insight into social organization at the Fujia site. Human Biology 87(1), 7184.Google Scholar
Downs, E.F. and Lowenstein, J.M. (1995) Identification of archaeological blood proteins: A cautionary note. Journal of Archaeological Science 22(1), 1116.Google Scholar
Dudd, S.N. and Evershed, R.P. (1998) Direct demonstration of milk as an element of archaeological economies. Science 282(5393), 1478–81.Google Scholar
Dudd, S.N., Evershed, R.P. and Gibson, A.M. (1999) Evidence for varying patterns of exploitation of animal products in different prehistoric pottery traditions based on lipids preserved in surface and absorbed residues. Journal of Archaeological Science 26(12), 1473–82.Google Scholar
Dungait, J.A.J., Doherty, G., Straker, V. and Evershed, R.P. (2010) Seasonal variations in bulk tissue, fatty acid and monosaccharide δ13C values of leaves from mesotrophic grassland plant communities under different grazing managements. Phytochemistry 71, 415–28.Google Scholar
Dungait, J.A., Stear, N.A., van Dongen, B.E., Bol, R. and Evershed, R.P. (2008) Off-line pyrolysis and compound-specific stable carbon isotope analysis of lignin moieties: a new method for determining the fate of lignin residues in soil. Rapid Communications in Mass Spectrometry 22(11), 1631–9.Google Scholar
Dunne, J., Evershed, R.P., Salque, M., Cramp, L., Bruni, S., Ryan, K., et al. (2012) First dairying in green Saharan Africa in the fifth millennium bc. Nature 486(7403), 390–4.Google Scholar
Dunne, J., Mercuri, A.M., Evershed, R.P., Bruni, S. and di Lernia, S. (2016) Earliest direct evidence of plant processing in prehistoric Saharan pottery. Nature Plants 3(1), 16194.Google Scholar
Dupras, T.L. and Tocheri, M.W. (2007) Reconstructing infant weaning histories at Roman period Kellis, Egypt using stable isotope analysis of dentition. American Journal of Physical Anthropology 134(1), 6374.Google Scholar
Dyson-Hudson, R. and Smith, E.A. (1978) Human territoriality: an ecological reassessment. American Anthropologist 80(1), 2141.Google Scholar
Ebersbach, R. (2002) Von Bauern und Rindern. Eine Ökosystemanalyse zur Bedeutung der Rinderhaltung in bäuerlichen Gesellschaften als Grundlage zur Modellbildung im Neolithikum. Basler Beiträge zur Archäologie 15.Google Scholar
Ebersbach, R., Ruckstuhl, B. and Bleicher, N. (2015) Zürich ‘Mozartstrasse’, Band 5. Die neolithischen Befunde und die Dendroarchäologie. Zürich und Egg: Fotorotar.Google Scholar
Ebersbach, R. and Schade, C. (2004) Modelling the intensity of Linear Pottery land use – an example from the Mörlener Bucht in the Wetterau Basin, Hesse, Germany. In: Wien, S. (ed.), Enter the Past – the E-way into the Four Dimensions of Cultural Heritage. Oxford: British Archaeological Reports, pp. 259–73.Google Scholar
Edwards, C.J., Bollongino, R., Scheu, A., Chamberlain, A., Tresset, A., Vigne, J.D., et al. (2007) Mitochondrial DNA analysis shows a Near Eastern Neolithic origin for domestic cattle and no indication of domestication of European aurochs. Proceedings of the Royal Society of London B: Biological Sciences 274(1616), 1377–85.Google Scholar
Eerkens, J.W. and Bartelink, E.J. (2013) Sex-biased weaning and early childhood diet among middle Holocene hunter-gatherers in Central California. American Journal of Physical Anthropology 152(4), 471–83.Google Scholar
Eerkens, J.W., Berget, A.G. and Bartelink, E.J. (2011) Estimating weaning and early childhood diet from serial micro-samples of dentin collagen. Journal of Archaeological Science 38(11), 3101–11.Google Scholar
Ellis, M.A., Merritt, C.W., Novak, S.A. and Dixon, K.J. (2011) The signature of starvation: a comparison of bone processing at a Chinese encampment in Montana and the Donner Party Camp in California. Historical Archaeology 45(2), 97112.Google Scholar
Emlen, J.M. (1966) The role of time and energy in food preference. American Naturalist 100, 611–17.Google Scholar
Emswiller, E. (2006) Genetic data and plant domestication. In: Zeder, M.A., Bradley, D.G., Emschwiller, E. and Smith, B.D. (eds), Documenting Domestication: New Genetic and Archaeological Paradigms. Berkeley: University of California Press, pp. 99122.Google Scholar
English Heritage (2008) Management of Research Projects in the Historic Environment: PPN 3: Archaeological Excavation. London: English Heritage.Google Scholar
Enk, J.M., Devault, A.M., Kuch, M., Murgha, Y.E., Rouillard, J.M. and Poinar, H.N. (2014) Ancient whole genome enrichment using baits built from modern DNA. Molecular Biology and Evolution 31(5), 1292–4.Google Scholar
Erasmus, U. (1986) Fat and Oils: the Complete Guide to Fats and Oils in Health and Nutrition. Vancouver: Alive Books.Google Scholar
Eren, M.I. and Outram, A.K. (2012) Preface to ‘Faunal Extinctions and Introductions’. World Archaeology 44(1), 12.Google Scholar
Ericson, J.E. (1985) Strontium isotope characterization in the study of prehistoric human ecology. Journal of Human Evolution 14(5), 503–14.Google Scholar
Ervynck, A., Lentacker, A., Müldner, G., Richards, M. and Dobney, K. (2007) An investigation into the transition from forest dwelling pigs to farm animals in medieval Flanders, Belgium. In: Albarella, U., Dobney, K., Ervynck, A. and Rowley-Conwy, P. (eds), Pigs and Humans, 10,000 Years of Interaction. Oxford: Oxford University Press, pp. 171–93.Google Scholar
Evans, J.A., Pashley, V., Richards, G.J., Brereton, N. and Knowles, T.G. (2015) Geogenic lead isotope signatures from meat products in Great Britain: Potential for use in food authentication and supply chain traceability. Science of the Total Environment 537, 447–52.Google Scholar
Evershed, R.P. (2008a) Organic residue analysis in archaeology: the archaeological biomarker revolution. Archaeometry 50(6), 895924.Google Scholar
Evershed, R.P. (2008b) Experimental approaches to the interpretation of absorbed organic residues in archaeological ceramics. World Archaeology 40(1), 2647.Google Scholar
Evershed, R.P., Arnot, K.I., Collister, J., Eglinton, G. and Charters, S. (1994) Application of isotope ratio monitoring gas chromatography–mass spectrometry to the analysis of organic residues of archaeological origin. Analyst 119(5), 909–14.Google Scholar
Evershed, R.P., Bethell, P.H., Reynolds, P.J. and Walsh, N.J. (1997b) 5β-Stigmastanol and related 5β-stanols as biomarkers of manuring: analysis of modern experimental material and assessment of the archaeological potential. Journal of Archaeological Science 24(6), 485–95.Google Scholar
Evershed, R.P., Dudd, S.N., Charters, S., Mottram, H.R., Stott, A.W., Raven, A., et al. (1999) Lipids as carriers of anthropogenic signals from prehistory. Philosophical Transactions of the Royal Society 354, 1931.Google Scholar
Evershed, R.P., Dudd, S.N., Lockheart, M.J. and Jim, S. (2001) Lipids in archaeology. In: Brothwell, D.R. and Pollard, A.M. (eds), Handbook of Archaeological Sciences. Oxford: Wiley, pp. 331–49.Google Scholar
Evershed, R.P., Heron, C. and Goad, L.J. (1991) Epicuticular wax components preserved in potsherds as chemical indicators of leafy vegetables in ancient diets. Antiquity 65, 540–4.Google Scholar
Evershed, R.P., Jerman, K. and Eglinton, G. (1985) Pine wood origin for pitch from the Mary Rose. Nature 314, 528–30.Google Scholar
Evershed, R.P., Payne, S., Sherratt, A.G., Copley, M.S., Coolidge, J., Urem-Kotsu, D., et al. (2008) Earliest date for milk use in the Near East and southeastern Europe linked to cattle herding. Nature 455, 528–31.Google Scholar
Evershed, R.P., Van Bergen, P.F., Peakman, T.M., Leigh-Firbank, E.C., Horton, M.C., Edwards, D., et al. (1997a) Archaeological frankincense. Nature 390(6661), 667–8.Google Scholar
Evershed, R.P., Vaughan, S.J., Dudd, S.N. and Soles, J.S. (1997c) Fuel for thought? Beeswax in lamps and conical cups from Late Minoan Crete. Antiquity 71, 979–85.Google Scholar
Ewbank, J.M., Phillipson, D.W., Whitehouse, R.D. and Higgs, E.S. (1964) Sheep in the Iron Age: a method of study. Proceedings of the Prehistoric Society 30, 423–6. Cambridge: Cambridge University Press.Google Scholar
Faegri, K. and Iversen, J. (1950) Textbook of Modern Pollen Analysis. Copenhagen: Einar Munksgaard.Google Scholar
Fagan, B. (2001) Graham Clark: An Intellectual Biography of an Archaeologist. Boulder, co: Westview Press.Google Scholar
Farquhar, G.D., Ehleringer, J.R. and Hubick, K.T. (1989) Carbon isotope discrimination and photosynthesis. Annual Review of Plant Physiology and Plant Molecular Biology 40: 503–37.Google Scholar
Fernandes, R., Nadeau, M.J. and Grootes, P.M. (2012) Macronutrient-based model for dietary carbon routing in bone collagen and bioapatite. Archaeological and Anthropological Sciences 4(4), 291301.Google Scholar
Ferrio, J.P., Araus, J.L., Buxó, R., Voltas, J. and Bort, J. (2005) Water management practices and climate in ancient agriculture: inferences from the stable isotope composition of archaeobotanical remains. Vegetation History and Archaeobotany 14(4), 510–17.Google Scholar
Feuerbach, L. (1850) Die Naturwissenschaft und die Revolution. Ludwig Feuerbach. Werke. Bd. 4, Kritiken und Abhandlungen III, pp. 243–65.Google Scholar
Feyerabend, P. (1975) Against Method: An Outline of an Anarchist Theory of Knowledge. London: NLB.Google Scholar
Fiedel, S. (1996) Blood from stones? Some methodological and interpretive problems in blood residue analysis. Journal of Archaeological Science 23(1), 139–47.Google Scholar
Flannery, K.V. (1969) Origins and ecological effects of early domestication in Iran and the Near East. In: Ucko, P.J. and Dimbleby, G.W. (eds), The Domestication and Exploitation of Plants and Animals. Chicago, il: Aldine, pp. 73100.Google Scholar
Flannery, K.V. (1976a) Empirical determination of catchments in Oaxaca and Tehuacan. In: Flannery, K.V. (ed.), The Early Mesoamerican Village. New York: Academic Press, pp. 103–17.Google Scholar
Flannery, K.V. (1976b) The village and its catchment area. In: Flannery, K.V. (ed.), The Early Mesoamerican Village. New York: Academic Press, pp. 163–87.Google Scholar
Foley, R. (1985) Optimality theory in anthropology. Man 20(2), 222–42.Google Scholar
Forbes, H. (1989) Of grandfathers and grand theories: the hierarchized ordering of responses to hazard in a Greek rural community. In: Halstead, P. and O’Shea, J. (eds), Bad Year Economics: Cultural Responses to Risk and Uncertainty. Cambridge: Cambridge University Press, pp. 8797.Google Scholar
Forde, C.D. (1934) Habitat, Economy and Society: A Geographical Introduction to Ethnology. London: Methuen.Google Scholar
Fournié, G., Pfeiffer, D.U. and Bendrey, R. (2017) Early animal farming and zoonotic disease dynamics: modelling brucellosis transmission in Neolithic goat populations. Open Science 4(2), 160943.Google Scholar
Frachetti, M.D. (2008) Pastoralist Landscapes and Social Interaction in Bronze Age Eurasia. Berkeley: University of California Press.Google Scholar
Frachetti, M.D. (2009) Differentiated landscapes and non-uniform complexity among Bronze Age societies of the Eurasian steppe. In: Hanks, B.K., Linduff, K.M. (eds), Social Complexity in Prehistoric Eurasia. Cambridge: Cambridge University Press, pp. 1946.Google Scholar
Frachetti, M. and Benecke, N. (2009) From sheep to (some) horses: 4500 years of herd structure at the pastoralist settlement of Begash (south-eastern Kazakhstan). Antiquity 83(322), 1023–37.Google Scholar
Frachetti, M.D., Spengler, R.N., Fritz, G.J. and Mar’yashev, A.N. (2010) Earliest direct evidence for broomcorn millet and wheat in the central Eurasian steppe region. Antiquity 84(326), 9931010.Google Scholar
France, D.L. (2008) Human and Nonhuman Bone Identification: A Color Atlas. Boca Raton, fl: CRC Press.Google Scholar
Fraser, F.C. and King, J.E. (1954) Faunal remains. In Clark, J.D.G. Excavations at Star Carr: An Early Mesolithic Site at Seamer Near Scarborough, Yorkshire. Cambridge: Cambridge University Press, pp. 7095.Google Scholar
Fraser, R., Bogaard, A., Charles, M., Styring, A.K., Wallace, M., Jones, G., et al. Ditchfield, (2013a) Assessing natural variation and the effects of charring, burial and pre-treatment on the stable carbon and nitrogen isotope values of archaeobotanical cereal and pulse remains. Journal of Archaeological Science 40: 4754–66.Google Scholar
Fraser, R.A., Bogaard, A., Heaton, T., Charles, M., Jones, G., Christensen, B.T., et al. (2011) Manuring and stable nitrogen isotope ratios in cereals and pulses: towards a new archaeobotanical approach to the inference of land use and dietary practices. Journal of Archaeological Science 38(10), 2790–804.Google Scholar
Fraser, R.A., Bogaard, A., Schäfer, M., Arbogast, R.-M. and Heaton, T.H.E.H. (2013b) Integrating botanical, faunal and human stable carbon and nitrogen isotope values to reconstruct land use and palaeodiet at LBK Vaihingen an der Enz, Baden-Württemberg. World Archaeology 45, 492517.Google Scholar
Frei, K.M., Coutu, A.N., Smiarowski, K., Harrison, R., Madsen, C.K., Arneborg, J., et al. (2015). Was it for walrus? Viking Age settlement and medieval walrus ivory trade in Iceland and Greenland. World Archaeology 47(3), 439466.Google Scholar
French, C. and Kousoulakou, M. (2003) Geomorphological and micromorphological investigations of palaeosols, valley sediments and a sunken floored dwelling at Botai, Kazakhstan. In: Levine, M.A., Renfrew, C. and Boyle, K.V. (eds), Prehistoric Steppe Adaptation and the Horse. Cambridge: McDonald Institute, pp. 105–14.Google Scholar
French, D.H., Hillman, G.C., Payne, S. and Payne, R.J. (1972) Excavations at Can Hasan III 1969–1970. In: Higgs, E.S. (ed.), Papers in Economic Prehistory: Studies by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press, pp. 181–90.Google Scholar
Fridrich, C. (1994) Kulturgeschichtliche Betrachtungen zur Bandkeramik im Merzbachtal. In: Lüning, J. and Stehli, P. (eds), Die Bandkeramik im Merzbachtal auf der Aldenhovener Platte. Cologne: Rheinland-Verlag GmbH, pp. 207393.Google Scholar
Frison, G.C. and Rehrer, C.A. (1970) Age determination of buffalo teeth eruption and wear. In: Frison, G.C. (ed.), Glenrock Buffalo Jump, 48CO304: Late Prehistoric Period Buffalo Procurement and Butchery. Plains Anthropologist Memoirs 7, pp. 4650.Google Scholar
Fry, E., Kim, S.K., Chigurapti, S., Mika, K.M., Ratan, A., Dammermann, A., et al. (2017) Accumulation and functional architecture of deleterious genetic variants during the extinction of Wrangel Island mammoths. bioRxiv, 137455.Google Scholar
Fullagar, R., Field, J., Denham, T. and Lentfer, C. (2006) Early and mid Holocene tool-use and processing of taro (Colocasia esculenta), yam (Dioscorea sp.) and other plants at Kuk Swamp in the highlands of Papua New Guinea. Journal of Archaeological Science 33(5), 595614.Google Scholar
Fuller, B.T., Fuller, J.L., Sage, N.E., Harris, D.A., O’Connell, T.C. and Hedges, R.E. (2005) Nitrogen balance and δ15N: why you’re not what you eat during nutritional stress. Rapid Communications in Mass Spectrometry 19(18), 2497–506.Google Scholar
Fuller, D.Q., Qin, L., Zheng, Y., Zhao, Z., Chen, X., Hosoya, L.A. and Sun, G.-P. (2009) The domestication process and domestication rate in rice: spikelet bases from the Lower Yangtze. Science 323, 1607–10.Google Scholar
Gallego Llorente, M.G., Jones, E.R., Eriksson, A., Siska, V., Arthur, K.W., Arthur, J.W., et al. (2015) Ancient Ethiopian genome reveals extensive Eurasian admixture in Eastern Africa. Science 350(6262), 820–2.Google Scholar
Gamba, C., Hanghøj, K., Gaunitz, C., Alfarhan, A.H., Alquraishi, S.A., Al-Rasheid, K.A., et al. (2015) Comparing the performance of three ancient DNA extraction methods for high-throughput sequencing. Molecular Ecology Resources. DOI: 10.1111/1755-0998.12470.Google Scholar
Gamble, C. (1979) Hunting strategies in the Central European Palaeolithic. In: Proceedings of the Prehistoric Society 45, 3552.Google Scholar
Gamkrelidze, T.V. and Ivanov, V. (1995) Indo-European and the Indo-Europeans: A Reconstruction and Historical Analysis of a Proto-Language and a Proto-Culture, Vol. 1. Berlin: Mouton de Gruyter.Google Scholar
García-Granero, J.J. (2015) A tale of multi-proxies: integrating macro- and microbotanical remains to understand subsistence strategies. Vegetation History and Archaeobotany 24, 121–33.Google Scholar
Gaunitz, C., Fages, A., Hanghøj, K., Albrechtsen, A., Khan, N., Schubert, M., et al. (2018) Ancient genomes revisit the ancestry of domestic and Przewalski’s horses. Science 360, 111–14.Google Scholar
Gerbault, P., Gillis, R., Vigne, J.-D., Tresset, A., Bréhard, S. and Thomas, M.G. (2016) Statistically robust representation and comparison of mortality profiles in archaeozoology. Journal of Archaeological Science 71, 2432.Google Scholar
Gerbault, P., Liebert, A., Itan, Y., Powell, A., Currat, M., Burger, J., et al. (2011) Evolution of lactase persistence: an example of human niche construction. Philosophical Transactions of the Royal Society of London B: Biological Sciences 366(1566), 863–77.Google Scholar
Gerling, C., Doppler, T., Heyd, V., Knipper, C., Kuhn, T., Lehmann, M.F., et al. (2017) High-resolution isotopic evidence of specialised cattle dairying in the European Neolithic. PLoS ONE 12: e0180164.Google Scholar
Germanov, P.G. and Kosintsev, P.A. (1995) Kostnye ostatki poseleniia pozdney bronzy Druzhnyy 1 v Iuzhnom Zaural’e, konf., Rossiia i Vostok: problemy vzaimodeystviia, v. 5, N 2., Cheliabinsk, pp. 116–18.Google Scholar
Germonpré, M., Sablin, M.V., Stevens, R.E., Hedges, R.E., Hofreiter, M., Stiller, M. and Després, V.R. (2009) Fossil dogs and wolves from Palaeolithic sites in Belgium, the Ukraine and Russia: osteometry, ancient DNA and stable isotopes. Journal of Archaeological Science 36(2), 473–90.Google Scholar
Gernaey, A.M., Waite, E.R., Collins, M.J., Craig, O.E. and Sokol, R.J. (2001) Survival and interpretation of archaeological protein. In: Brothwell, D.R. and Pollard, A.M. (eds), Handbook of Archaeological Sciences. Chichester: Wiley, pp. 323–9.Google Scholar
Gilbert, B.M. (1990) Mammalian Osteology. Columbia: Missouri Archaeological Society.Google Scholar
Giles, R.J. and Brown, T.A. (2008) Improved methodology for extraction and amplification of DNA from single grains of charred wheat. Journal of Archaeological Science 35(9), 2585–8.Google Scholar
Gillis, R., Bréhard, S., Bălăşescu, A., Ughetto-Monfrin, J., Popovici, D., Vigne, J.D. and Balasse, M. (2013) Sophisticated cattle dairy husbandry at Borduşani-Popină (Romania, fifth millennium bc): the evidence from complementary analysis of mortality profiles and stable isotopes. World Archaeology 45(3), 447–72.Google Scholar
Gillis, R.E., Kovačiková, L., Bréhard, S., Guthmann, E., Vostrovská, I., Nohálová, H., et al. (2017). The evolution of dual meat and milk cattle husbandry in Linearbandkeramik societies. Proceedings of the Royal Society B: Biological Sciences 284, 20170905.Google Scholar
Glassow, M.A. (1978) The concept of carrying capacity in the study of cultural process. Advances in Archaeological Method and Theory 1, 3148.Google Scholar
Gokhman, D., Lavi, E., Prüfer, K., Fraga, M.F., Riancho, J.A., Kelso, J., et al. (2014) Reconstructing the DNA methylation maps of the Neandertal and the Denisovan. Science 344(6183), 523–7.Google Scholar
Gokhman, D., Malul, A. and Carmel, L. (2017) Inferring past environments from ancient epigenomes. Molecular Biology and Evolution, p.msx211.Google Scholar
Gokhman, D., Meshorer, E. and Carmel, L. (2016) Epigenetics: it’s getting old. Past meets future in paleoepigenetics. Trends in Ecology & Evolution 31(4), 290300.Google Scholar
Gott, B., Barton, H., Samuel, D. and Torrence, R. (2006) Biology of starch. In: Torrence, R. and Barton, H. (eds), Ancient Starch Research. Walnut Creek, ca: Left Coast Press, pp. 3545.Google Scholar
Graham, R.W., Belmecheri, S., Choy, K., Culleton, B.J., Davies, L.J., Froese, D., et al. (2016) Timing and causes of mid-Holocene mammoth extinction on St. Paul Island, Alaska. Proceedings of the National Academy of Sciences 113(33), 9310–14.Google Scholar
Green, R.E., Krause, J., Ptak, S.E., Briggs, A.W., Ronan, M.T., Simons, J.F., et al. (2006) Analysis of one million base pairs of Neanderthal DNA. Nature 444(7117), 330–6.Google Scholar
Green, R.E., Malaspinas, A.S., Krause, J., Briggs, A.W., Johnson, P.L., Uhler, C., et al. (2008) A complete Neandertal mitochondrial genome sequence determined by high-throughput sequencing. Cell 134(3), 416–26.Google Scholar
Greenfield, H.J. (1999) The origins of metallurgy: distinguishing stone from metal cut-marks on bones from archaeological sites. Journal of Archaeological Science 26(7), 797808.Google Scholar
Gregoricka, L.A. and Judd, M.A. (2016) Isotopic evidence for diet among historic Bedouin of Khirbat al-Mudayna, Jordan. International Journal of Osteoarchaeology 26(4), 705–15.Google Scholar
Grigg, D. (1979) Ester Boserup’s theory of agrarian change: a critical review. Progress in Human Geography 3, 6484.Google Scholar
Grigson, C. (1981) Fauna. In: Simmons, I. and Tooley, M. (eds), The Environment in British Prehistory. London: Duckworth, pp. 110–24.Google Scholar
Gross, B.L. and Zhao, Z. (2014) Archaeological and genetic insights into the origins of domesticated rice. Proceedings of the National Academy of Sciences 111(17), 6190–7.Google Scholar
Grupe, G. and McGlynn, G.C. (2016) Preface. In: Grupe, G. and McGlynn, G.C. (eds), Isotopic Landscapes in Bioarchaeology. Heidleberg: Springer, pp. vviii.Google Scholar
Gu, Y., Zhao, Z. and Pearsall, D.M. (2013) Phytolith morphology research on wild and domesticated rice species in East Asia. Quaternary International 287, 141–8.Google Scholar
Guiry, E.J., Hepburn, J.C. and Richards, M.P. (2016) High-resolution serial sampling for nitrogen stable isotope analysis of archaeological mammal teeth. Journal of Archaeological Science 69, 21–8.Google Scholar
Gunn, A. (2009). Essential Forensic Biology, 2nd edn. Chichester: John Wiley & Sons.Google Scholar
Haak, W., Brandt, G., de Jong, H.N., Meyer, C., Ganslmeier, R., Heyd, V., et al. (2008) Ancient DNA, Strontium isotopes, and osteological analyses shed light on social and kinship organization of the Later Stone Age. Proceedings of the National Academy of Sciences 105(47), 18226–31.Google Scholar
Haak, W., Forster, P., Bramanti, B., Matsumura, S., Brandt, G., Tänzer, M., et al. (2005) Ancient DNA from the first European farmers in 7500-year-old Neolithic sites. Science 310(5750), 1016–18.Google Scholar
Haak, W., Lazaridis, I., Patterson, N., Rohland, N., Mallick, S., Llamas, B., et al. (2015) Massive migration from the steppe was a source for Indo-European languages in Europe. Nature 522, 207–11.Google Scholar
Hachem, L. (1999) Apport de l’archéozoologie a la connaissance de l’organisation villageoise rubanée. In: Breamer, F., Cleuziou, S. and Coudart, A. (eds), Habitat et société, XIXe Rencontres d’Archéologie et d’Histoire d’Antibes. Antibes: APDCA, pp. 325–38.Google Scholar
Hachem, L. (2000) New observations on the Bandkeramik house and social organization. Antiquity 74, 308–12.Google Scholar
Hafner, A. (2013) UNESCO World Heritage ‘prehistoric pile-dwellings around the Alps’: chances and challenges for management and research of cultural heritage under water. In Roio, M. (ed.), The Changing Coastal and Maritime Culture. The 5th Baltic Sea Region Cultural Heritage Forum, Tallinn 18–20 September 2013. Tallin: Estonian National Heritage Board.Google Scholar
Haile, J., Froese, D.G., MacPhee, R.D., Roberts, R.G., Arnold, L.J., Reyes, A.V., et al. (2009) Ancient DNA reveals late survival of mammoth and horse in interior Alaska. Proceedings of the National Academy of Sciences 106(52), 22352–7.Google Scholar
Hajdinjak, M., Fu, Q., Hübner, A., Petr, M., Mafessoni, F., Grote, S., et al. (2018) Reconstructing the genetic history of late Neanderthals. Nature 555(7698), 652–6.Google Scholar
Halstead, P. (1989a) Like rising damp? An ecological approach to the spread of farming in south east and central Europe. In: Milles, A., Williams, D. and Gardner, N. (eds), The Beginnings of Agriculture. Oxford: British Archaeological Reports, pp. 2353.Google Scholar
Halstead, P. (1989b) The economy has a normal surplus: economic stability and social change among early farming communities. In: Halstead, P. and O’Shea, J. (eds), Bad Year Economics: Cultural Responses to Risk and Uncertainty. Cambridge: Cambridge University Press, pp. 6880.Google Scholar
Halstead, P. (1992) From reciprocity to redistribution: modelling the exchange of livestock in Neolithic Greece. Anthropozoologica 16, 1930.Google Scholar
Halstead, P. (2003) Texts and bones: contrasting Linear B and archaeozoological evidence for animal exploitation in Mycenaean southern Greece. British School at Athens Studies 9, 257–61.Google Scholar
Halstead, P. (2014) Two Oxen Ahead: Pre-Mechanized Farming in the Mediterranean. Chichester: Wiley Blackwell.Google Scholar
Halstead, P. and Isaakidou, V. (2011) A pig fed by hand is worth two in the bush: ethnoarchaeology of pig husbandry in Greece and its archaeological implications. In: Albarella, U. (ed.), Ethnozooarchaeology: the Present Past of Human–Animal Relationships. Oxford: Oxbow, pp. 160–74.Google Scholar
Halstead, P. and O’Shea, J. (eds) (1989a) Bad Year Economics: Cultural Responses to Risk and Uncertainty. Cambridge: Cambridge University Press.Google Scholar
Halstead, P. and O’Shea, J. (1989b) Introduction: cultural responses to risk and uncertainty. In: Halstead, P. and O’Shea, J. (eds), Bad Year Economics: Cultural Responses to Risk and Uncertainty. Cambridge: Cambridge University Press, pp. 17.Google Scholar
Hamilton, J. and Thomas, R. (2012) Pannage, pulses and pigs: isotopic and zooarchaeological evidence for changing pig management practices in later medieval England. Medieval Archaeology 56(1), 234–59.Google Scholar
Hanghøj, K., Seguin-Orlando, A., Schubert, M., Madsen, T., Pedersen, J.S., Willerslev, E. and Orlando, L. (2016) Fast, accurate and automatic ancient nucleosome and methylation maps with epiPALEOMIX. Molecular Biology and Evolution 33(12), 3284–98.Google Scholar
Hanks, B.K. and Linduff, K.M. (2009) Reconsidering steppe social complexity within world prehistory. In: Hanks, B.K. and Linduff, K.M. (eds), Social Complexity in Prehistoric Eurasia. Cambridge: Cambridge University Press, pp. 17.Google Scholar
Hansel, F.A., Copley, M.S., Madureira, L.A. and Evershed, R.P. (2004) Thermally produced ω-(o-alkylphenyl) alkanoic acids provide evidence for the processing of marine products in archaeological pottery vessels. Tetrahedron Letters 45(14), 29993002.Google Scholar
Harcourt, R.A. (1971) Animal bones from Durrington Walls. In: Wainwright, G.J. and Longworth, I.H. (eds), Durrington Walls Excavations 1966–1968. London: Society of Antiquaries, pp. 338–50.Google Scholar
Harding, R.M., Fullerton, S.M., Griffiths, R.C., Bond, J., Cox, M.J., Schneider, J.A., et al. (1997) Archaic African and Asian lineages in the genetic ancestry of modern humans. American Journal of Human Genetics 60(4), 772–89.Google Scholar
Harris, D.R. and Hillman, G.C. (eds) (1989) Foraging and Farming: The Evolution of Plant Exploitation. London: Unwin Hyman.Google Scholar
Hart, J.P., Urquhart, G.R., Feranec, R.S. and Lovis, W.A. (2009) Non-linear relationship between bulk δ13C and percent maize in carbonized cooking residues and the potential of false-negatives in detecting maize. Journal of Archaeological Science 36(10), 2206–12.Google Scholar
Hart, R.H. (2001) Where the buffalo roamed – or did they? Great Plains Research 11(1), 83102.Google Scholar
Haruda, A.F. (2014) Central Asian economies and ecologies in the Late Bronze Age: geometric morphometrics of the caprid astragalus and zooarchaeological investigations of pastoralism. PhD thesis, University of Exeter. http://hdl.handle.net/10871/17496.Google Scholar
Haruda, A.F. (2017) Separating sheep (Ovis aries L.) and goats (Capra hircus L.) using geometric morphometric methods: an investigation of astragalus morphology from late and final Bronze Age central Asian contexts. International Journal of Osteoarchaeology 27(4), 551–62.Google Scholar
Haslam, M. (2004) The decomposition of starch grains in soils: implications for archaeological residue analysis. Journal of Archaeological Science 31, 1715–34.Google Scholar
Hastorf, C.A. (2009) Rio Balsas most likely region for maize domestication. Proceedings of the National Academy of Sciences 106(13), 4957–8.Google Scholar
Hastorf, C.A. and Popper, V.S. (eds) (1988) Current Paleoethnobotany: Analytical Methods and Cultural Interpretations of Archaeological Plant Remains. Chicago, il: University of Chicago Press.Google Scholar
Hayden, B. (1975) The carrying capacity dilemma: An alternate approach. Memoirs of the Society for American Archaeology 30, 1121.Google Scholar
Healy, K. (2017) Fuck nuance. Sociological Theory 35(2), 118–27.Google Scholar
Heard, E. and Martienssen, R.A. (2014) Transgenerational epigenetic inheritance: myths and mechanisms. Cell 157(1), 95109.Google Scholar
Heck, L., Wilson, L.A., Evin, A., Stange, M. and Sánchez-Villagra, M.R. (2018) Shape variation and modularity of skull and teeth in domesticated horses and wild equids. Frontiers in Zoology 15(1), 14.Google Scholar
Hedges, S.B. and Schweitzer, M.H. (1995) Detecting dinosaur DNA. Science 268(5214), 1191–2.Google Scholar
Helbaek, H. (1959) The domestication of food plants in the Old World. Science 130, 365–72.Google Scholar
Hendy, J., Colonese, A.C., Franz, I., Fernandes, R., Fischer, R., Orton, D., et al. (2018a) Ancient proteins from ceramic vessels at Çatalhöyük West reveal the hidden cuisine of early farmers. Nature Communications 9(1), 4064.Google Scholar
Hendy, J., Warinner, C., Bouwman, A., Collins, M.J., Fiddyment, S., Fischer, R., et al. (2018b) Proteomic evidence of dietary sources in ancient dental calculus. Proceedings of the Royal Society B 285(1883), 20180977.Google Scholar
Hendy, J., Welker, F., Demarchi, B., Speller, C., Warinner, C. and Collins, M.J. (2018c) A guide to ancient protein studies. Nature Ecology and Evolution 2, 791–9.Google Scholar
Henry, A.G., Hudson, H.F. and Piperno, D.R. (2009) Changes in starch grain morphologies from cooking. Journal of Archaeological Science 36, 915–22.Google Scholar
Henry, D.O. (1985) Preagricultural sedentism: the Natufian example. In: Price, T.D. and Brown, J.A. (eds), Prehistoric Hunter-Gatherers: The Emergence of Cultural Complexity. Orlando, fl: Academic Press, pp. 365–84.Google Scholar
Henry, D.O. (1989) From Foraging to Agriculture. Philadelphia: University of Pennsylvania Press.Google Scholar
Herring, D.A., Saunders, S.R. and Katzenberg, M.A. (1998) Investigating the weaning process in past populations. American Journal of Physical Anthropology 105(4), 42539.Google Scholar
Hershkovitz, I., Donoghue, H.D., Minnikin, D.E., May, H., Lee, O.Y.C., Feldman, M., et al. (2015) Tuberculosis origin: the Neolithic scenario. Tuberculosis 95, S122S126.Google Scholar
Higgs, E.S. (ed.) (1972) Papers in Economic Prehistory: Studies by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press.Google Scholar
Higgs, E.S. (ed.) (1975) Palaeoeconomy: Being the Second Volume of Papers in Economic Prehistory by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press.Google Scholar
Higgs, E.S. and Jarman, M.R. (1975) Palaeoeconomy. In: Higgs, E.S. (ed.), Palaeoeconomy: Being the Second Volume of Papers in Economic Prehistory by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press, pp. 18.Google Scholar
Higgs, E.S. and Vita-Finzi, C. (1972) Prehistoric economies: a territorial approach. In: Higgs, E.S. (ed.), Papers in Economic Prehistory: Studies by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press, pp. 2736.Google Scholar
Higham, C. (1966) Stock rearing in prehistoric Europe, with special reference to the Danish islands and the Alpine foreland. Unpublished thesis: University of Cambridge.Google Scholar
Higham, C.F. (1968a) Stock rearing as a cultural factor in prehistoric Europe. Proceedings of the Prehistoric Society 33, 84106.Google Scholar
Higham, C.F.W. (1968b) Patterns of prehistoric economic exploitation on the Alpine Foreland. Vierteljahrsschrift der Naturforschenden Gesellschaft in Zürich 113, 4192.Google Scholar
Higham, C.F.W. (1969) Towards an economic prehistory of Europe. Current Anthropology 10(2/3), 139–50.Google Scholar
Higham, C.F.W. and Message, M.A. (1969) An assessment of a prehistoric technique of bovine husbandry. In: Brothwell, D. and Higgs, E. (eds), Science in Archaeology. London: Thames & Hudson, pp. 315–30.Google Scholar
Higham, T.G., Jacobi, R.M. and Ramsey, C.B. (2006) AMS radiocarbon dating of ancient bone using ultrafiltration. Radiocarbon 48(2), 179–95.Google Scholar
Higuchi, R., Bowman, B., Freiberger, M., Ryder, O.A. and Wilson, A.C. (1984) DNA sequences from the quagga, an extinct member of the horse family. Nature 312, 282–4.Google Scholar
Hilditch, T.P. and Williams, P.N. (1964) The Chemical Composition of Natural Fats, 4th edn. New York: Wiley.Google Scholar
Hill, J.D. (1995) Ritual and Rubbish in the Iron Age of Wessex. Oxford: British Archaeological Reports.Google Scholar
Hillman, G. (1981) Reconstructing crop husbandry practices from charred remains of crops. In: Mercer, R. (ed.), Farming Practice in British Prehistory. Edinburgh: Edinburgh University Press, pp. 123–62.Google Scholar
Hillman, G. (1984) Interpretation of archaeological plant remains: the application of ethnographic models from Turkey. In: Van Zeist, W. and Casperie, W.A. (eds), Plants and Ancient Man. Rotterdam: Balkema, pp. 141.Google Scholar
Hillman, G. (1996) Late Pleistocene changes in wild plant-foods available to hunter-gatherers of the northern Fertile Crescent: possible preludes to cereal cultivation. In: Harris, D.R. (ed.), The Origins and Spread of Agriculture and Pastoralism in Eurasia. Washington, dc: Smithsonian Press.Google Scholar
Hillman, G.C. and Davies, M.S. (1990a) Measured domestication rates in wild wheats and barley under primitive cultivation, and their archaeological implications. Journal of World Prehistory 4(2), 157222.Google Scholar
Hillman, G.C. and Davies, M.S. (1990b) Domestication rates in wild-type wheats and barley under primitive cultivation. Biological Journal of the Linnean Society 39(1), 3978.Google Scholar
Hodder, I. (1986) Reading the Past: Current Approaches to Interpretation in Archaeology. Cambridge: Cambridge University Press.Google Scholar
Hodder, I. (1990) The Domestication of Europe. Oxford: Blackwell.Google Scholar
Hodder, I. (2012) Introduction: contemporary theoretical debate in archaeology. In Hodder, I. (ed.) Archaeological Theory Today, 2nd edn. Cambridge: Polity Press, pp. 114.Google Scholar
Hodder, I. and Orton, C. (1976) Spatial Analysis in Archaeology. Cambridge: Cambridge University Press.Google Scholar
Hodgson, J.A. and Disotell, T.R. (2008) No evidence of a Neanderthal contribution to modern human diversity. Genome Biology 9(2), 206.Google Scholar
Hodson, M.J., Parker, A.G., Leng, M.J. and Sloane, H.J. (2008) Silicon, oxygen and carbon isotope composition of wheat (Triticum aestivum L.) phytoliths: implications for palaeoecology and archaeology. Journal of Quaternary Science 23, 331–9.Google Scholar
Holl, H.M., Brooks, S.A., Archer, S., Brown, K., Malvick, J., Penedo, M.C.T. and Bellone, R.R. (2016) Variant in the RFWD3 gene associated with PATN1, a modifier of leopard complex spotting. Animal Genetics 47(1), 91101.Google Scholar
Hong, C., Jiang, H., , E., Wu, Y., Guo, L., Xie, Y., et al. (2012) Identification of milk component in ancient food residue by proteomics. PloS one 7(5), e37053.Google Scholar
Hosch, S. and Jacomet, S. (2004) Ackerbau und Sammelwirtschaft. Ergebnisse der Untersuchung von Samen und Früchten. In: Jacomet, S., Leuzinger, U. and Schibler, J. (eds), Die jungsteinzeitliche Seeufersiedlung Arbon Bleiche 3: Umwelt und Wirtschaft. Frauenfeld: Amt für Archäologie des Kantons Thurgau, pp. 112–57.Google Scholar
Howland, M.R., Corr, L.T., Young, S.M., Jones, V., Jim, S., Van Der Merwe, N.J., et al. (2003) Expression of the dietary isotope signal in the compound-specific δ13C values of pig bone lipids and amino acids. International Journal of Osteoarchaeology 13(1–2), 5465.Google Scholar
Hu, Y., Hu, S., Wang, W., Wu, X., Marshall, F. B., Chen, X., et al. (2014) Earliest evidence for commensal processes of cat domestication. Proceedings of the National Academy of Sciences 111(1), 116–20.Google Scholar
Hullar, M.A. and Fu, B.C. (2014) Diet, the gut microbiome, and epigenetics. Cancer Journal 20(3), 170–5.Google Scholar
Hunt, H.V., Vander Linden, M., Liu, X., Motuzaite-Matuzevicuite, G. and Jones, M.K. (2008) Millets across Eurasia: chronology and context of early records of the genera Panicum and Setaria from archaeological sites in the Old World. Vegetation History and Archaeobotany 17, S5–S18.Google Scholar
Hurcombe, L.M. (2014) Perishable Material Culture in Prehistory: Investigating the Missing Majority. Abingdon: Routledge.Google Scholar
Ingold, T. (1980) Hunter, Pastoralists and Ranchers. Cambridge: Cambridge University Press.Google Scholar
Iriarte, J. (2003) Assessing the feasibility of identifying maize through the analysis of cross-shaped size and three-dimensional morphology of phytoliths in the grasslands of southeastern South America. Journal of Archaeological Science 30, 1085–94.Google Scholar
Iriarte, J. (2007) New perspectives on early plant domestication and the development of agriculture in the Americas. In: Denham, T., Iriarte, J. and Vrydaghs, L. (eds) Rethinking Agriculture: Archaeological and Ethnoarchaeological Perspectives, Walnut Creek, ca: Left Coast Press, pp. 167–88.Google Scholar
Iriarte, J., Holst, I., Marozzi, O., Listopad, C., Alonso, E., Rinderknecht, A. and Montaña, J. (2004) Evidence for cultivar adoption and emerging complexity during the mid-Holocene in the La Plata basin. Nature 432, 614–17.Google Scholar
Iriarte, J., Power, M.J., Rostain, S., Mayle, F.E., Jones, H., Watling, J., et al. (2012) Fire-free land use in pre-1492 Amazonian savannas. Proceedings of the National Academy of Sciences 109, 6473–8.Google Scholar
Isaakidou, V. (2006) Ploughing with cows: Knossos and the secondary products revolution. In: Serjeantson, D. and Field, D. (eds), Animals in the Neolithic of Britain and Europe. Oxford: Oxbow Books, pp. 95112.Google Scholar
Isaakidou, V. (2011) Farming regimes in Neolithic Europe: gardening with cows and other models. In: Hadjikoumis, A., Robinson, E. and Viner, S. (eds), The Dynamics of Neolithisation in Europe: Studies in honour of Andrew Sherratt. Oxford: Oxbow, pp. 90112.Google Scholar
Itan, Y., Powell, A., Beaumont, M.A., Burger, J. and Thomas, M.G. (2009) The origins of lactase persistence in Europe. PLoS Computational Biology 5(8), p.e1000491.Google Scholar
Iversen, J. (1941) Landnam i Danmarks stenalder: En pollenanalytisk undersøgelse over det første landbrugs indvirkning paa vegetationsudviklingen. Danmarks Geologiske Undersogelse, Series 2, 66, 168.Google Scholar
Iversen, J. (1949) The influence of prehistoric man on vegetation. Danmarks Geologiske Undersogelse, Series 4, 3, 125.Google Scholar
Jackson, R. (1972) A vicious circle? The consequences of Von Thünen in tropical Africa. Area 4(4), 258–61.Google Scholar
Jacomet, S. and Brombacher, C. (2005) Reconstructing intra-site patterns in Neolithic lakeshore settlements: the state of archaeobotanical research and future prospects. In: Della Casa, P. and Trachsel, M. (eds) WES’04 – Wetland Economies and Societies. Proceedings of the International Conference in Zurich, 10–13 March 2004. Zurich: Chronos, pp. 6994.Google Scholar
Jacomet, S., Brombacher, C. and Dick, M. (1989) Archäobotanik am Zürichsee. Ackerbau, Sammelwirtschaft und Umwelt von neolithischen und bronzezeitlichen Seeufersiedlungen im Raum Zürich. Zurich: Orell Füssli Verlag.Google Scholar
Jacomet, S., Ebersbach, R., Akeret, Ö., Antolín, F., Baum, T., Bogaard, A., et al. (2016) On-site data cast doubts on the hypothesis of shifting cultivation in the late Neolithic (c. 4300–2400 cal bc): Landscape management as an alternative paradigm. The Holocene 26, 1858–74.Google Scholar
Jacomet, S., Leuzinger, U. and Schibler, J. (2004) Die jungsteinzeitliche Seeufersiedlung Arbon Bleiche 3: Umwelt und Wirtschaft, Archäologie im Thurgau. Frauenfeld: Amt für Archäologie des Kantons Thurgau.Google Scholar
Jaenicke-Despres, V., Buckler, E.S., Smith, B.D., Gilbert, M.T.P., Cooper, A., Doebley, J. and Pääbo, S. (2003) Early allelic selection in maize as revealed by ancient DNA. Science 302(5648), 1206–8.Google Scholar
Jakucs, J., Bánffy, E., Oross, K., Voicsek, V., Bronk Ramsey, C., Dunbar, E., et al. (2016) Between the Vinča and Linearbandkeramik worlds: the diversity of practices and identities in the 54th-53rd centuries cal bc in southwest Hungary and beyond. Journal of World Prehistory 29, 267336.Google Scholar
Janowitz Koch, I., Clark, M.M., Thompson, M.J., Deere-Machemer, K.A., Wang, J., Duarte, L., et al. (2016) The concerted impact of domestication and transposon insertions on methylation patterns between dogs and grey wolves. Molecular Ecology 25(8), 1838–55.Google Scholar
Jansen, T., Forster, P., Levine, M.A., Oelke, H., Hurles, M., Renfrew, C., Weber, J. and Olek, K. (2002) Mitichondrial DNA and the origins of the domestic horse. Proceedings of the National Academy of Sciences 99(16), 10905–10.Google Scholar
Jarman, H.N. (1972) The origins of wheat and barley cultivation. In: Higgs, E.S. (ed.), Papers in Economic Prehistory: Studies by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press, pp. 1526.Google Scholar
Jarman, H.N., Legge, A.J. and Charles, J.A. (1972) Retrieval of plant remains from archaeological sites by froth flotation. In: Higgs, E.S. (ed.), Papers in Economic Prehistory: Studies by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press, pp. 3948.Google Scholar
Jarman, M.R., Bailey, G.N. and Jarman, H.N. (eds) (1982) Early European Agriculture: Its Foundation and Development. Cambridge: Cambridge University Press.Google Scholar
Jarman, M.R., Vita-Finzi, C. and Higgs, E.S. (1972) Site catchment analysis in archaeology. In: Ucko, P.J., Tringham, R. and Dimbleby, G.W. (eds), Man, Settlement and Urbanism. London: Duckworth, pp. 61–6.Google Scholar
Jarman, M.R. and Webley, D. (1975) Settlement and land use in Capitanata, Italy. In: Higgs, E.S. (ed.), Palaeoeconomy: Being the Second Volume of Papers in Economic Prehistory by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press, pp. 177222.Google Scholar
Jay, M., Fuller, B.T., Richards, M.P., Knüsel, C.J. and King, S.S. (2008) Iron Age breastfeeding practices in Britain: isotopic evidence from Wetwang Slack, East Yorkshire. American Journal of Physical Anthropology 136(3), 327–37.Google Scholar
Jenkins, E. (2009) Phytolith taphonomy: a comparison of dry ashing and acid extraction on the breakdown of conjoined phytoliths formed in Triticum durum. Journal of Archaeological Science 36, 2402–7.Google Scholar
Jenkins, E., Jamjoum, K., Nuimat, S., Stafford, R., Nortcliff, S. and Mithen, S. (2016) Identifying ancient water availability through phytolith analysis: an experimental approach. Journal of Archaeological Science 73, 8293.Google Scholar
Jensen, P. (2014) Behaviour epigenetics–The connection between environment, stress and welfare. Applied Animal Behaviour Science 157, 17.Google Scholar
Jeong, C., Wilkin, S., Amgalantugs, T., Bouwman, A.S., Taylor, W.T.T., Hagan, R.W., et al. (2018) Bronze Age population dynamics and the rise of dairy pastoralism on the eastern Eurasian steppe. Proceedings of the National Academy of Sciences 115(48), E11248–E11255.Google Scholar
Johnson, E.V., Parmenter, P.C. and Outram, A.K. (2016) A new approach to profiling taphonomic history through bone fracture analysis, with an example application to the Linearbandkeramik site of Ludwinowo 7. Journal of Archaeological Science: Reports 9, 623–9.Google Scholar
Johnson, E.V., Timpson, A., Thomas, M.G. and Outram, A.K. (2018) Reduced intensity of bone fat exploitation correlates with increased potential access to dairy fats in early Neolithic Europe. Journal of Archaeological Science 94, 60–9.Google Scholar
Johnson, M. (1999) Archaeological Theory: An Introduction. Oxford: Blackwell.Google Scholar
Jones, B.A., Grace, D., Kock, R., Alonso, S., Rushton, J., Said, M.Y., et al. (2013) Zoonosis emergence linked to agricultural intensification and environmental change. Proceedings of the National Academy of Sciences 110(21), 8399–404.Google Scholar
Jones, E.R., Zarina, G., Moiseyev, V., Lightfoot, E., Nigst, P.R., Manica, A., et al. (2017) The Neolithic transition in the Baltic was not driven by admixture with early European farmers. Current Biology 27(4), 576–82.Google Scholar
Jones, G. (1984) Interpretation of archaeological plant remains: ethnographic models from Greece. In Van Zeist, W. and Casperie, W.A. (eds) Plants and Ancient Man. Rotterdam: Balkema, pp. 4361.Google Scholar
Jones, G., Bogaard, A., Halstead, P., Charles, M. and Smith, H. (1999) Identifying the intensity of crop husbandry practices on the basis of weed floras. Annals of the British School at Athens 94, 167–89.Google Scholar
Jones, G., Bogaard, A., Charles, M. and Hodgson, J.G. (2000a) Distinguishing the effects of agricultural practices relating to fertility and disturbance: a functional ecological approach in archaeobotany. Journal of Archaeological Science 27, 1073–84.Google Scholar
Jones, G., Valamoti, S. and Charles, M. (2000b) Early crop diversity: a ‘new’ glume wheat from northern Greece. Vegetation History and Archaeobotany 9, 133–46.Google Scholar
Jones, M., Hunt, H., Lightfoot, E., Lister, D., Liu, X. and Motuzaite-Matuzeviciute, G. (2011) Food globalization in prehistory. World Archaeology 43(4), 665–75.Google Scholar
Judkins, G., Smith, M. and Keys, E. (2008) Determinism within human–environment research and the rediscovery of environmental causation. The Geographical Journal 174(1), 1729.Google Scholar
Julien, M.A., Bocherens, H., Burke, A., Drucker, D.G., Patou-Mathis, M., Krotova, O. and Péan, S. (2012) Were European steppe bison migratory? 18O, 13C and Sr intra-tooth isotopic variations applied to a palaeoethological reconstruction. Quaternary International 271, 106–19.Google Scholar
Kadwell, M., Fernandez, M., Stanley, H.F., Baldi, R., Wheeler, J.C., Rosadio, R. and Bruford, M.W. (2001) Genetic analysis reveals the wild ancestors of the llama and the alpaca. Proceedings of the Royal Society of London B: Biological Sciences 268(1485), 2575–84.Google Scholar
Kafil, H.S., baha Hosseini, S., Sohrabi, M. and Asgharzadeh, M. (2014) Brucellosis: presence of zoonosis infection 3,500 years ago in north of Iran. Asian Pacific Journal of Tropical Disease 4, S684S686.Google Scholar
Kalieva, S.S. and Logvin, V.N. (1997) Skotovody Turgaya v Tret’em Tysyacheletii do Nashej Ehry, Kustanai: Kustanai University.Google Scholar
Karr, L.P., Outram, A.K. and Hannus, L.A. (2010) A chronology of bone marrow and bone grease exploitation at the Mitchell Prehistoric Indian Village. Plains Anthropologist 55(215), 215–23.Google Scholar
Karr, L.P., Short, E.G., Hannus, L.A. and Outram, A.K. (2015) A bone grease processing station at the Mitchell Prehistoric Indian Village: Archaeological evidence for the exploitation of bone fats. Environmental Archaeology 20(1), 112Google Scholar
Kasparov, A.K. and Outram, A.K. (2013) Nekotorye zamechaniia po povodu osteologicheskih materialov kompleksa Lisakovky (poselenie i mogil’niki epohi bronzy). In: Usmanova, E. (ed.), Pamiatniki Lisakovskoy okrugi: arheologicheskie siuzhety. Karaganda: Tengri, pp. 230–9.Google Scholar
Katzenberg, M.A., Herring, D.A. and Saunders, S.R. (1996) Weaning and infant mortality: evaluating the skeletal evidence. American Journal of Physical Anthropology 101(S23), 17799.Google Scholar
Kealhofer, L. (2002) Changing perceptions of risk: the development of agro-ecosystems in Southeast Asia. American Anthropologist 104, 178–94.Google Scholar
Keeley, L.H. (1988) Hunter-gatherer economic complexity and ‘population pressure’: A cross-cultural analysis. Journal of Anthropological Archaeology 7(4), 373411.Google Scholar
Keeley, L.H. (1991) Ethnographic models for Late Glacial hunter-gatherers. In: Barton, N., Roberts, A.J. and Roe, D.A. (eds), The Late Glacial in North-West Europe: Human Adaptation and Environmental Change at the End of the Pleistocene. York: Council for British Archaeology, pp. 179–90.Google Scholar
Keene, A. (1979) Economic optimization models and the study of hunter-gatherer subsistence settlement systems. In: Renfrew, C. and Cooke, K. (eds), Transformations: Mathematical Approaches to Culture Change. New York: Academic Press, pp. 369404.Google Scholar
Kelekna, P. (2009) The Horse in Human History. Cambridge: Cambridge University Press.Google Scholar
Kelly, E.F., Amundson, R.G., Marino, B.D. and DeNiro, M.J. (1991) Stable isotope ratios of carbon in phytoliths as a quantitative method of monitoring vegetation and climate change. Quaternary Research 35, 222–33.Google Scholar
Kelly, R.L. (1995) The Foraging Spectrum: Diversity in Hunter-Gatherer Lifeways. Washington, dc: Smithsonian Institution Press.Google Scholar
Kelly, R.L. (2013) The Lifeways of Hunter-Gatherers: The Foraging Spectrum. Cambridge: Cambridge University Press.Google Scholar
Kendall, I.P., Lee, M.R. and Evershed, R.P. (2017) The effect of trophic level on individual amino acid δ15N values in a terrestrial ruminant food web. STAR: Science & Technology of Archaeological Research 3(1), 135–45.Google Scholar
Kim, J. (2017) Gut microbiome, a potent modulator of epigenetics in human diseases. Journal of Bacteriology and Virology 47(2), 7586.Google Scholar
Kislenko, A. and Tatarintseva, N. (1999) The eastern Ural steppe at the end of the Stone Age. In: Levine, M., Rassamakin, Y., Kislenko, A. and Tatarintseva, N. (eds), Late Prehistoric Exploitation of the Eurasian Steppe. Cambridge: McDonald Institute, pp. 183216.Google Scholar
Klein, R.G. and Cruz-Uribe, K. (1984) The Analysis of Animal Bones from Archaeological Sites. Chicago, il: University of Chicago Press.Google Scholar
Knipper, C. (2011). Die räumliche Organisation der linearbandkeramischen Rinderhaltung: naturwissenschaftliche und archäologische Untersuchungen. BAR International Series 2035. Oxford: Archaeopress.Google Scholar
Knörzer, K.-H. (1971). Urgeschichtliche Unkräuter im Rheinland: ein Beitrag zur Entstehungsgeschichte der Segetalgesellschaften. Vegetatio 23, 89111.Google Scholar
Knudson, K.J., Williams, H.M., Buikstra, J.E., Tomczak, P.D., Gordon, G.W. and Anbar, A.D. (2010) Introducing δ88Sr/86Sr analysis in archaeology: a demonstration of the utility of strontium isotope fractionation in paleodietary studies. Journal of Archaeological Science 37, 2352–64.Google Scholar
Koch, P.L., Fisher, D.C. and Dettman, D. (1989) Oxygen isotope variation in the tusks of extinct proboscideans: a measure of season of death and seasonality. Geology 17(6), 515–19.Google Scholar
Koch, P.L., Tuross, N. and Fogel, M.L. (1997) The effects of sample treatment and diagenesis on the isotopic integrity of carbonate in biogenic hydroxylapatite. Journal of Archaeological Science 24(5), 417–29.Google Scholar
Kohl, P.L. (2007) The Making of Bronze Age Eurasia. Cambridge: Cambridge University Press.Google Scholar
Kohler, T.A., Smith, M.E., Bogaard, A., Feinman, G.M., Peterson, C.E., Betzenhauser, A., et al. (2017) Greater post-Neolithic wealth disparities in Eurasia than in North America and Mesoamerica. Nature 551(7682), 619–22.Google Scholar
Kohn, M.J. (1999) You are what you eat. Science 283(5400), 335–6.Google Scholar
Kooyman, B., Newman, M. E. and Ceri, H. (1992) Verifying the reliability of blood residue analysis on archaeological tools. Journal of Archaeological Science 19(3), 265–9.Google Scholar
Koryakova, L. and Epimakhov, A. (2007) The Urals and Western Siberia in the Bronze and Iron Ages. Cambridge: Cambridge University Press.Google Scholar
Kosiba, S.B., Tykot, R.H. and Carlsson, D. (2007) Stable isotopes as indicators of change in the food procurement and food preference of Viking Age and Early Christian populations on Gotland (Sweden). Journal of Anthropological Archaeology 26(3), 394411.Google Scholar
Kosintsev, P.A. (1989) Ohota i skotovodstvo u naseleniia lesostepnogo Zaural’ia v epohu bronzy, Stanovlenie i razvitie proizvodiashego hoziaystva na Urale, UrO AN SSSR, Sverdlovsk, pp. 84108.Google Scholar
Krause-Kyora, B., Makarewicz, C., Evin, A., Flink, L.G., Dobney, K., Larson, G., et al. (2013) Use of domesticated pigs by Mesolithic hunter-gatherers in northwestern Europe. Nature Communications 4, 2348.Google Scholar
Kremenetski, C.V., Tarasov, P.E. and Cherkinsky, A.E. (1997) Postglacial development of Kazakhstan pine forests. Geographie physique et Quaternaire 51(3), 391404.Google Scholar
Kreuz, A. (1990) Die ersten Bauern Mitteleuropas – eine archäobotanische Untersuchung zur Umwelt und Landwirtschaft der Ältesten Bandkeramik. Analecta Praehistorica Leidensia 23. Leiden: University of Leiden.Google Scholar
Kreuz, A. (2007) Archaeobotanical perspectives on the beginning of agriculture north of the Alps. In: Colledge, S. and Conolly, J. (eds), The Origins and Spread of Domestic Plant in Southwest Asia and Europe. Walnut Creek, ca: Left Coast Press, pp. 259–94.Google Scholar
Kreuz, A. (2012) Die Vertreibung aus dem Paradies? Archäobiologische Ergebnisse zum Frühneolithikum im westlichen Mitteleuropa. Bericht der Römisch-Germanischen Kommission 91, 23196.Google Scholar
Kreuz, A and Marinova, E. (2017) Archaeobotanical evidence of crop growing and diet within the areas of the Karanovo and the Linear Pottery Cultures: a quantitative and qualitative approach. Vegetation History and Archaeobotany 26, 639–57.Google Scholar
Kreuz, A., Marinova, E.M., Schäfer, E. and Wiethold, J. (2005) A comparison of early Neolithic crop and weed assemblages from the Linearbandkeramik and the Bulgarian Neolithic cultures: differences and similarities. Vegetation History and Archaeobotany 14, 237–58.Google Scholar
Kruk, J. (1973) Studia Osadnicze nad Neolitem Wyzyn Lessowych (Studies on the Neolithic Settlement of the Loess Uplands). Warsaw: Polska Akademia Nauk, Instytut Historii Kultury Materialnej.Google Scholar
Kuhn, S.L. and Miller, D.S. (2015) Artifacts as patches: the marginal value theorem and stone tool life histories. In: Goodale, N. and Andrefsky, W., Jr (eds), Lithic Technological Systems and Evolutionary Theory. Cambridge: Cambridge University Press, pp. 172–97.Google Scholar
Kuhn, T.S. (1996) The Structure of Scientific Revolutions. 3rd edn. Chicago, il: University of Chicago Press.Google Scholar
Kuzmina, I.E. (1997) Horses of North Eurasia from the Pliocene Till the Present Time. St. Petersburg: Zoological Institute of the Russian Academy of Sciences.Google Scholar
Laland, K.N. and O’Brien, M.J. (2010) Niche Construction Theory and archaeology. Journal of Archaeological Method and Theory 17(4), 303–22.Google Scholar
Lang, C., Peters, J., Pöllath, N., Schmidt, K. and Grupe, G. (2013) Gazelle behaviour and human presence at early Neolithic Göbekli Tepe, south-east Anatolia. World Archaeology 45(3), 410–29.Google Scholar
Lantos, I., Spangenberg, J.E., Giovannetti, M.A., Ratto, N. and Maier, M.S. (2015) Maize consumption in pre-Hispanic south-central Andes: chemical and microscopic evidence from organic residues in archaeological pottery from western Tinogasta (Catamarca, Argentina). Journal of Archaeological Science 55, 8399.Google Scholar
Larson, G., Albarella, U., Dobney, K., Rowley-Conwy, P., Schibler, J., Tresset, A., et al. (2007) Ancient DNA, pig domestication, and the spread of the Neolithic into Europe. Proceedings of the National Academy of Sciences 104(39), 15276–81.Google Scholar
Larson, G. and Burger, J. (2013) A population genetics view of animal domestication. Trends in Genetics 29, 195205.Google Scholar
Larson, G., Dobney, K., Albarella, U., Fang, M., Matisoo-Smith, E., Robins, J., Lowden, S., et al. (2005) Worldwide phylogeography of wild boar reveals multiple centers of pig domestication. Science 307(5715), 1618–21.Google Scholar
Larson, G. and Fuller, D.Q. (2014) The evolution of animal domestication. Annual Review of Ecology, Evolution, and Systematics 45, 115–36.Google Scholar
Lee, E.S., Song, E.J. and Nam, Y.D. (2017) Dysbiosis of gut microbiome and its impact on epigenetic regulation. Journal of Clinical Epigenetics 3(S1), 14.Google Scholar
Lee, R.B. and DeVore, I. (1968) Problems in the study of hunters and gatherers. In: Lee, R.B. and DeVore, I. (eds), Man the Hunter. Chicago, il: Aldine, pp. 312.Google Scholar
Legge, A.J. (1972) Prehistoric exploitation on gazelle in Palestine. In: Higgs, E.S. (ed.), Palaeoeconomy: Being the Second Volume of Papers in Economic Prehistory by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press, pp. 119–24.Google Scholar
Legge, A.J. (1981) Aspects of cattle husbandry. In: Mercer, R. (ed.), Farming Practice in British Prehistory. Edinburgh: Edinburgh University Press, pp. 169–81.Google Scholar
Legge, A.J. (2005) Milk use in prehistory: the osteological evidence. In: Mulville, J. and Outram, A. K. (eds), The Zooarchaeology of Fats, Oils, Milk and Dairying. Oxford: Oxbow, pp. 813.Google Scholar
Legge, A.J. and Rowley-Conwy, P.A. (1988) Star Carr Revisited. London: Centre for Extra Mural Studies.Google Scholar
Le Huray, J.D. and Schutkowski, H. (2005) Diet and social status during the La Tene period in Bohemia: carbon and nitrogen stable isotope analysis of bone collagen from Kutná Hora-Karlov and Radovesice. Journal of Anthropological Archaeology 24(2), 135–47.Google Scholar
Leonardi, M., Boschin, F., Giampoudakis, K., Beyer, R.M., Krapp, M., Bendrey, R., et al. (2018) Late Quaternary horses in Eurasia in the face of climate and vegetation change. Science Advances 4(7), eaar5589.Google Scholar
Leonardi, M., Gerbault, P., Thomas, M.G. and Burger, J. (2012) The evolution of lactase persistence in Europe. A synthesis of archaeological and genetic evidence. International Dairy Journal 22(2), pp. 8897.Google Scholar
Leonardi, M., Librado, P., Der Sarkissian, C., Schubert, M., Alfarhan, A.H., Alquraishi, S.A., et al. (2017) Evolutionary patterns and processes: lessons from ancient DNA. Systematic Biology 66(1), e1–e29.Google Scholar
Lepetz, S. (2013) Horse sacrifice in a Pazyryk culture kurgan: the princely tomb of Berel’ (Kazakhstan). Selection criteria and slaughter procedures. Anthropozoologica 48(2), 309–21.Google Scholar
Leuzinger, U. (2002) Holzartefakte. In: de Capitani, A., Deschler-Erb, S., Leuzinger, U., Marti-Grädel, E. and Schibler, J. (eds), Die jungsteinzeitliche Siedlung Arbon Bleiche 3: Funde. Frauenfeld: Amt für Archäologie des Kantons Thurgau, pp. 76114.Google Scholar
Levine, M. (1999) The origins of horse husbandry on the Eurasian steppe. In: Levine, M., Rassamakin, Y., Kislenko, A. and Tatarintseva, N. (eds), Late Prehistoric Exploitation of the Eurasian Steppe. Cambridge: McDonald Institute. pp. 558.Google Scholar
Levine, M. (2004) Exploring the criteria for early horse domestication. In: Jones, M. (ed.), Traces of Ancestry: Studies in Honour of Colin Renfrew. Cambridge: McDonald Institute. pp. 115–26.Google Scholar
Levine, M.A. (2005) Domestication and early history of the horse. In: Mills, D. and McDonnell, S. (eds), The Domestic Horse: The Origins, Development and Management of its Behavior. Cambridge: Cambridge University Press, pp. 522.Google Scholar
Levine, M.A., Bailey, G.N., Whitwell, K.E. and Jeffcott, L.B. (2000) Palaeopathology and horse domestication. In: Bailey, G., Charles, R. and Winder, N. (eds), Human Ecodynamics and Environmental Archaeology. Oxford: Oxbow, pp. 123–33.Google Scholar
Levine, M. and Kislenko, A.M. (1997) New Eneolithic and Early Bronze Age radiocarbon dates for north Kazakhstan and south Siberia. Cambridge Archaeological Journal 7(2), 297300.Google Scholar
Levine, M.A., Whitwell, K.E. and Jeffcott, L.B. (2002) A Romano-British horse burial from Icklingham, Suffolk. Archaeofauna 11, 63102.Google Scholar
Levine, M.A., Whitwell, K.E. and Jeffcott, L.B. (2005) Abnormal thoracic vertebrae and the evolution of horse husbandry. Archaeofauna: International Journal of Archaeozoology 14, 93109.Google Scholar
Lewis, J., Pike, A.W.G., Coath, C.D. and Evershed, R.P. (2017) Strontium concentration, radiogenic (87Sr/86Sr) and stable (δ88Sr) strontium isotope systematics in a controlled feeding study. Science and Technology of Archaeological Research 3(1), 5365.Google Scholar
Lewontin, R.C. (1982) Organism and environment. In: Plotkin, H.C. (ed.), Learning, Development and Culture. New York: Wiley, pp. 151–70.Google Scholar
Lewontin, R.C. (1983) Gene, organism and environment. In: Bendall, D.S. (ed.), Evolution from Molecules to Men. Cambridge: Cambridge University Press, pp. 273–85.Google Scholar
Liamputtong, P. (2007) Childrearing and Infant Care Issues: A cross-cultural perspective. New York: Nova Publishers.Google Scholar
Libby, W.F. (1946) Atmospheric helium three and radiocarbon from cosmic radiation. Physical Review 69(11–12), 671.Google Scholar
Librado, P., Der Sarkissian, C., Ermini, L., Schubert, M., Jónsson, H., Albrechtsen, A., et al. (2015) Tracking the origins of Yakutian horses and the genetic basis for their fast adaptation to subarctic environments. Proceedings of the National Academy of Sciences 112(50), E6889–E6897.Google Scholar
Librado, P., Gamba, C., Gaunitz, C., Der Sarkissian, C., Pruvost, M., Albrechtsen, A., et al. (2017) Ancient genomic changes associated with domestication of the horse. Science 356(6336), 442–5.Google Scholar
Lidén, K., Eriksson, G., Nordqvist, B., Götherström, A. and Bendixen, E. (2004) ‘The wet and the wild followed by the dry and the tame’–or did they occur at the same time? Diet in Mesolithic–Neolithic southern Sweden. Antiquity 78(299), 2333.Google Scholar
Liebert, A., Lopez, S., Jones, B.L., Montalva, N., Gerbault, P., Lau, W., et al. (2017) World-wide distributions of lactase persistence alleles and the complex effects of recombination and selection. Human Genetics 136, 1445–53.Google Scholar
Lightfoot, E., Liu, X. and Jones, M.K. (2013) Why move starchy cereals? A review of the isotopic evidence for prehistoric millet consumption across Eurasia. World Archaeology 45(4), 574623.Google Scholar
Lightfoot, E., Motuzaite-Matuzeviciute, G., O’Connell, T.C., Kukushkin, I.A., Loman, V., Varfolomeev, V., et al. (2015) How ‘pastoral’ is pastoralism? Dietary diversity in Bronze Age communities in the Central Kazakhstan Steppes. Archaeometry 57 (Suppl. 1), 232–49.Google Scholar
Lipson, M. et al. (2017) Parallel palaeogenomic trancests reveal complex genetic history of early European farmers. Nature 551, 368–72.Google Scholar
Linderholm, A. (2015) Ancient DNA: the next generation–chapter and verse. Biological Journal of the Linnean Society 117(1), 150–60.Google Scholar
Linderholm, A., Jonson, C. H., Svensk, O. and Lidén, K. (2008) Diet and status in Birka: stable isotopes and grave goods compared. Antiquity 82(316), 446–61.Google Scholar
Linderholm, A. and Larson, G. (2013) The role of humans in facilitating and sustaining coat colour variation in domestic animals. Seminars in Cell & Developmental Biology 24(6), 587–93.Google Scholar
Lindgren, G., Backström, N., Swinburne, J., Hellborg, L., Einarsson, A., Sandberg, K., Cothran, G., et al. (2004) Limited number of patrilines in horse domestication. Nature Genetics 36(4), 335–6.Google Scholar
Lindqvist, C. and Possnert, G. (1997) The subsistence economy and diet at Jakobs/Ajvide, Eksta parish and other prehistoric dwelling and burial sites on Gotland in long-term perspective. In: Burenhult, G. (ed.), Remote sensing 1, Stockholm: Stockholm University, pp. 2990.Google Scholar
Llamas, B., Valverde, G., Fehren-Schmitz, L., Weyrich, L.S., Cooper, A. and 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 & Technology of Archaeological Research 3(1), 114.Google Scholar
Longinelli, A. (1984) Oxygen isotopes in mammal bone phosphate: a new tool for paleohydrological and paleoclimatological research? Geochimica et Cosmochimica Acta 48(2), 385–90.Google Scholar
Loy, T.H. (1983) Prehistoric blood residues: detection on tool surfaces and identification of species of origin. Science 220, 1269–71.Google Scholar
Lubell, D., Jackes, M., Schwarcz, H., Knyf, M. and Meiklejohn, C. (1994) The Mesolithic–Neolithic transition in Portugal: isotopic and dental evidence of diet. Journal of Archaeological Science 21(2), 201–16.Google Scholar
Ludwig, A., Pruvost, M., Reissmann, M., Benecke, N., Brockmann, G. A., Castaños, P., et al. (2009) Coat color variation at the beginning of horse domestication. Science 324(5926), 485.Google Scholar
Ludwig, A., Reissmann, M., Benecke, N., Bellone, R., Sandoval-Castellanos, E., Cieslak, M., et al. (2015) Twenty-five thousand years of fluctuating selection on leopard complex spotting and congenital night blindness in horses. Philosophical Transactions of the Royal Society B 370(1660), 20130386.Google Scholar
Lüning, J. (2000) Steinzeitliche Bauern in Deutschland – die Landwirtschaft im Neolithikum. Universitätsforschungen zur prähistorischen Archäologie aus dem Seminar für Vor- und Frühgeschichte der Universität Frankfurt/M, vol 58. Bonn: Dr. Rudolf Habelt GmbH.Google Scholar
Lüning, J. (2006) Missionare aus dem Western bekehren und belehren. Archäologie in Deutschland 2006(3), 2831.Google Scholar
Lyman, R.L. (1994) Vertebrate Taphonomy. Cambridge: Cambridge University Press.Google Scholar
Lyman, R.L. (2008) Quantitative Paleozoology. Cambridge: Cambridge University Press.Google Scholar
Lyotard, J.-F. (1984) The Postmodern Condition: A Report on Knowledge. Manchester: Manchester University Press.Google Scholar
MacArthur, R.H. and Pianka, E.R. (1966) On optimal use of a patchy environment. American Naturalist 100, 603–9.Google Scholar
MacHugh, D.E., Larson, G. and Orlando, L. (2017) Taming the past: ancient DNA and the study of animal domestication. Annual Review of Animal Biosciences 5, 329–51.Google Scholar
MacHugh, D.E., Shriver, M.D., Loftus, R.T., Cunningham, P. and Bradley, D.G. (1997) Microsatellite DNA variation and the evolution, domestication and phylogeography of taurine and zebu cattle (Bos taurus and Bos indicus). Genetics 146(3), 1071–86.Google Scholar
MacNeish, R.S. and Eubanks, M.W. (2000) Comparative analysis of the Río Balsas and Tehuacán models for the origin of maize. Latin American Antiquity 11, 320.Google Scholar
Madella, M. and Lancelotti, C. (2012) Taphonomy and phytoliths: A user manual. Quaternary International 275, 7683.Google Scholar
Madella, M., García-Granero, J.J., Out, W.A., Ryan, P. and Usai, D. (2014) Microbotanical evidence of domestic cereals in Africa 7000 years ago. PLoS One 9(10), e110177.Google Scholar
Madella, M., Jones, M.K., Echlin, P., Powers-Jones, A. and Moore, M. (2009) Plant water availability and analytical microscope of phytoliths: Implications for ancient irrigation in arid zones. Quaternary International 193, 3240.Google Scholar
Madsen, D.B. and Schmitt, D.N. (1998) Mass collecting and the diet breadth model: A Great Basin example. Journal of Archaeological Science 25(5), 445–55.Google Scholar
Mallory, J.P. (1996) In Search of the Indo-Europeans: Language, Archaeology and Myth. London: Thames & Hudson.Google Scholar
Maier, U. and Vogt, R. (2001) Botanische und pedologische Untersuchungen zur Ufersiedlung Hornstaad-Hörnle IA. Stuttgart: Konrad Theiss Verlag.Google Scholar
Makarewicz, C.A. (2018) Stable isotopes in pastoralist archaeology as indicators of diet, mobility, and animal husbandry practices. In: Ventresca Miller, A.R. and Makarewicz, C.A. (eds), Isotopic Investigations of Pastoralism in Prehistory. Abingdon: Routledge, pp. 141–58.Google Scholar
Makarewicz, C.A. and Sealy, J. (2015) Dietary reconstruction, mobility, and the analysis of ancient skeletal tissues: expanding the prospects of stable isotope research in archaeology. Journal of Archaeological Science 56, 146–58.Google Scholar
Makarova, L.A. (1976) Harakteristika kostnogo materiala iz poseleniia Sargary, Proshloe Kazakhstana po arkheologicheskim istochnikam, Almaty, pp. 211–26.Google Scholar
Makarova, L.A. (1977) Zhivotnye Atasu i drugih poseleniy Central’nogo Kazakhstana, Arkheologicheskie isledovaniia v Otrare, Almaty, pp. 124–31.Google Scholar
Makarova, L.A. (1980) Kosti zhivotnyh iz dvuh poseleniy epohi bronzy v Severnom Kazakhstane, Arkheologicheskie isledovaniia drevnego i srednevekovogo Kazakhstana, Almaty, pp. 141–51.Google Scholar
Marciniak, A. (2005) Placing Animals in the Neolithic: Social Zooarchaeology of Prehistoric Communities. London: UCL Press.Google Scholar
Marom, N. and Bar-Oz, G. (2009) Culling profiles: the indeterminacy of archaeozoological data to survivorship curve modelling of sheep and goat herd maintenance strategies. Journal of Archaeological Science 36(5), 1184–7.Google Scholar
Matisoo-Smith, E. and Horsbugh, K.A. (2012) DNA for Archaeologists. Walnut Creek, ca: Left Coast Press.Google Scholar
Matlova, V., Roffet-Salque, M., Pavlu, I., Kyselka, J., Sedlarova, I., Filip, V. and Evershed, R.P. (2017) Defining pottery use and animal management at the Neolithic site of Bylany (Czech Republic). Journal of Archaeological Science: Reports 14, 262–74.Google Scholar
Matsui, A. (2008) Fundamentals of Zooarchaeology in Japan. Kyoto: Kyoto University Press.Google Scholar
Matsuoka, Y., Vigouroux, Y., Goodman, M.M., Sanchez, J., Buckler, E. and Doebley, J. (2002) A single domestication for maize shown by multilocus microsatellite genotyping. Proceedings of the National Academy of Sciences 99(9), 6080–4.Google Scholar
Matuzeviciute, G.M., Lightfoot, E., O’Connell, T.C., Voyakin, D., Liu, X., Loman, V., et al. (2015) The extent of cereal cultivation among the Bronze Age to Turkic period societies of Kazakhstan determined using stable isotope analysis of bone collagen. Journal of Archaeological Science 59, 2334.Google Scholar
Mayr, C., Grupe, G., Toncala, A. and Lihl, C.M. (2016) Linking oxygen isotopes of animal phosphate with altimetry, results from archaeological finds from a transect in the Alps. In: Grupe, G. and McGlynn, G.V. (eds), Isotopic Landscapes in Bioarchaeology. Heidelberg: Springer, pp. 157–72.Google Scholar
McClaran, M.P. and Umlauf, M. (2000) Desert grassland dynamics estimated from carbon isotopes in grass phytoliths and soil organic matter. Journal of Vegetation Science 11, 71–6.Google Scholar
McCorriston, J. and Hole, F. (1991) The ecology of seasonal stress and the origins of agriculture in the Near East. American Anthropologist 93(1), 4669.Google Scholar
Mercader, J., Abtosway, M., Baquedano, E., Bird, R.W., Díez-Martín, F., Domínguez-Rodrigo, M., et al. (2017) Starch contamination landscapes in field archaeology: Olduvai Gorge, Tanzania. Boreas, DOI: 10.1111/bor.12241.Google Scholar
Mertz, I.V. and Mertz, V.K. (2013) Novye materialy rannego bronzovogo veka iz Zapadnoj chasti Kulundinskoj ravniny. In: Kubareva, G.A. and Semibratov, V.P. (eds), Sohranenie i izuchenie kul’turnogo nasledija Altajskogo kraja: materialy XVIII XIX regional’nyh nauchno-prakticheskih konferencij. Barnaul: AZBUKA, pp. 207–15.Google Scholar
Metcalfe, D. and Jones, K.T. (1988) A reconsideration of animal body part utility indices. American Antiquity 59(1), 486504.Google Scholar
Mileto, S., Kaiser, E., Rassamakin, Y. and Evershed, R.P. (2017) New insights into the subsistence economy of the Eneolithic Dereivka culture of the Ukrainian North-Pontic region through lipid residues analysis of pottery vessels. Journal of Archaeological Science: Reports 13, 6774.Google Scholar
Miller, G.L. (2015) Ritual economy and craft production in small-scale societies: evidence from microwear analysis of Hopewell bladelets. Journal of Anthropological Archaeology 39, 124–38.Google Scholar
Miller, N.F., Spengler, R.N. and Frachetti, M. (2016) Millet cultivation across Eurasia: origins, spread, and the influence of seasonal climate. The Holocene 26(10), 1566–75.Google Scholar
Milner, N., Craig, O.E., Bailey, G.N., Pedersen, K. and Andersen, S.H. (2004) Something fishy in the Neolithic? A re-evaluation of stable isotope analysis of Mesolithic and Neolithic coastal populations. Antiquity 78(299), 922.Google Scholar
Minagawa, M. and Wada, E. (1984) Stepwise enrichment of 15N along food chains: further evidence and the relation between δ15N and animal age. Geochimica et Cosmochimica Acta 48(5), 1135–40.Google Scholar
Miyake, Y. and Wada, E. (1967) The abundance ratio of 15N/14N in marine environments. Records of Oceanographic Works in Japan 9(1), 3753.Google Scholar
Molleson, T. and Blondiaux, J. (1994) Riders’ bones from Kish, Iraq. Cambridge Archaeological Journal 4(2), 312–16.Google Scholar
Montgomery, J., Beaumont, J., Jay, M., Keefe, K., Gledhill, A.R., Cook, G.T., et al. (2013a) Strategic and sporadic marine consumption at the onset of the Neolithic: increasing temporal resolution in the isotope evidence. Antiquity 87(338), 1060–72.Google Scholar
Montgomery, J., Budd, P. and Evans, J. (2000) Reconstructing the lifetime movements of ancient people: a Neolithic case study from southern England. European Journal of Archaeology 3(3), 370–85.Google Scholar
Montgomery, J., Evans, J.A., Chenery, S.R., Pashley, V. and Killgrove, K. (2010) ‘Gleaming, white and deadly’: using lead to track human exposure and geographic origins in the Roman period in Britain. Journal of Roman Archaeology; supplementary series 78, 199226.Google Scholar
Montgomery, J., Evans, J.A. and Horstwood, M.S. (2013b) Evidence for long-term averaging of strontium in bovine enamel using TIMS and LA-MC-ICP-MS strontium isotope intra-molar profiles. Environmental Archaeology 15(1), 3242.Google Scholar
Moore, A.M.T. and Hillman, G.C. (1992) The Pleistocene to Holocene transition and human economy in Southwest Asia: The impact of the Younger Dryas. American Antiquity, 482–94.Google Scholar
Moore, A.M.T., Hillman, G.C. and Legge, A.J. (1975) The excavation of Tell Abu Hureyra in Syria: a preliminary report. Proceedings of the Prehistoric Society 41, 5077.Google Scholar
Moore, A.M.T., Hillman, G.C. and Legge, A.J. (2000) Village on the Euphrates: From Foraging to Farming at Abu Hureyra. Oxford: Oxford University Press.Google Scholar
Morgan, E.D., Titus, L., Small, R.J. and Edwards, C. (1984) Gas-chromatographic analysis of fatty material from a Thule midden. Archaeometry 26, 43–8.Google Scholar
Morris, J. (2008) Re-examining associated bone groups from southern England and Yorkshire c. 4000 bc to ad 1550. Unpublished PhD thesis, University of Bournemouth.Google Scholar
Motuzaite-Matuzeviciute, G., Staff, R.A., Hunt, H.V., Liu, X. and Jones, M.K. (2013) The early chronology of broomcorn millet (Panicum miliaceum) in Europe. Antiquity 87(338), 1073–85.Google Scholar
Mukherjee, A.J., Copley, M.S., Berstan, R., Clark, K.A. and Evershed, R.P. (2005) Interpretation of δ13C values of fatty acids in relation to animal husbandry, food processing and consumption in prehistory. In: Mulville, J. and Outram, A.K. (eds), The Zooarchaeology of Fats, Oils, Milk and Dairying. Oxford: Oxbow, pp. 7793.Google Scholar
Mukherjee, A.J., Gibson, A.M. and Evershed, R.P. (2008) Trends in pig product processing at British Neolithic Grooved Ware sites traced through organic residues in potsherds. Journal of Archaeological Science 35(7), 2059–73.Google Scholar
Mullis, K.B. and Faloona, F.A. (1987) Specific synthesis of DNA in vitro via a polymerase-catalyzed chain reaction. Methods in Enzymology 155, 335–50.Google Scholar
Mullis, K., Faloona, F., Scharf, S., Saiki, R.K., Horn, G.T. and Erlich, H. (1986) Specific enzymatic amplification of DNA in vitro: the polymerase chain reaction. Cold Spring Harbor Symposia on Quantitative Biology 51, 263–73.Google Scholar
Murray, D.C., Haile, J., Dortch, J., White, N.E., Haouchar, D., Bellgard, M.I., et al. (2013) Scrapheap Challenge: A novel bulk-bone metabarcoding method to investigate ancient DNA in faunal assemblages. Scientific Reports 3, 337.Google Scholar
Mutolo, M.J., Jenny, L.L., Buszek, A.R., Fenton, T.W. and Foran, D.R. (2012) Osteological and molecular identification of brucellosis in ancient Butrint, Albania. American Journal of Physical Anthropology 147(2), 254–63.Google Scholar
Nagel, M.C. (1982) Frederick Soddy: From Alchemy to Isotopes. Journal of Chemical Education 59(9), 739–40.Google Scholar
Nätt, D., Rubin, C.J., Wright, D., Johnsson, M., Beltéky, J., Andersson, L. and Jensen, P. (2012) Heritable genome-wide variation of gene expression and promoter methylation between wild and domesticated chickens. BMC Genomics 13, 59.Google Scholar
Nehlich, O. (2015) The application of sulphur isotope analyses in archaeological research: a review. Earth-Science Reviews 142, 117.Google Scholar
Nelson, B.K., DeNiro, M.J., Schoeninger, M.J., De Paolo, D.J. and Hare, P.E. (1986) Effects of diagenesis on strontium, carbon, nitrogen and oxygen concentration and isotopic composition of bone. Geochimica et Cosmochimica Acta 50(9), 1941–9.Google Scholar
Nenquin, J. (1961) Salt: A Study in Economic Prehistory. Brugge: De Tempel.Google Scholar
Neil, S., Evans, J., Montgomery, J. and Scarre, C. (2016) Isotopic evidence for residential mobility of farming communities during the transition to agriculture in Britain. Royal Society Open Science 3(1), 150522.Google Scholar
Nistelberger, H.M., Smith, O., Wales, N., Star, B. and Boessenkool, S. (2016) The efficacy of high-throughput sequencing and target enrichment on charred archaeobotanical remains. Scientific Reports 6, 37347.Google Scholar
Nitsch, E., Andreou, S., Creuzieux, A., Gardeisen, A., Halstead, P., Isaakidou, V., et al. (2017) A bottom-up view of food surplus: using stable carbon and nitrogen isotope analysis to investigate agricultural strategies and diet at Bronze Age Archontiko and Thessaloniki Toumba, northern Greece. World Archaeology 49, 105–37.Google Scholar
Nitsch, E.K., Charles, M. and Bogaard, A. (2015) Calculating a statistically robust δ13C and δ15N offset for charred cereal and pulse seeds. STAR: Science & Technology of Archaeological Research 1(1), 18.Google Scholar
Nitsch, E.K., Humphrey, L.T. and Hedges, R.E. (2011) Using stable isotope analysis to examine the effect of economic change on breastfeeding practices in Spitalfields, London, UK. American Journal of Physical Anthropology 146(4), 619–28.Google Scholar
Noonan, J.P., Coop, G., Kudaravalli, S., Smith, D., Krause, J., Alessi, J., et al. (2006) Sequencing and analysis of Neanderthal genomic DNA. Science 314(5802), 1113–18.Google Scholar
O’Connell, T., Levine, M. and Hedges, R. (2003) The importance of fish in the diet of Central Eurasian peoples from the Mesolithic to the Early Iron Age. In: Levine, M., Renfrew, C. and Boyle, K. (eds), Prehistoric Steppe Adaptation and the Horse. Cambridge: McDonald Institute for Archaeological Research, pp. 253–68.Google Scholar
O’Connor, T. (2013) Animals as Neighbours: The Past and Present of Commensal Species. East Lansing: Michigan State University Press.Google Scholar
O’Day, S.J., van Neer, W. and Ervynck, A. (eds) (2004) Behaviour Behind Bones: The Zooarchaeology of Ritual, Religion, Status and Identity. Oxford: Oxbow Books.Google Scholar
Odum, E.P. (1959) Fundamentals of Ecology. London: Saunders.Google Scholar
Olsen, S.J. (1964) Mammal Remains from Archaeological Sites, Part 1: Southeastern and Southwestern United States. Cambridge, ma: Harvard University Press.Google Scholar
Olsen, S.L. (1989) Solutré: a theoretical approach to the reconstruction of upper Palaeolithic hunting strategies. Journal of Human Evolution 18(4), 295327.Google Scholar
Olsen, S. L. (1996) Horse hunters of the Ice Age. In: Olsen, S. (ed.), Horses through Time. Boulder, co: Roberts Rinehart, pp. 3556.Google Scholar
Olsen, S.L. (2003) The exploitation of horses at Botai, Kazakhstan. In: Levine, M., Renfrew, C. and Boyle, K. (eds), Prehistoric Steppe Adaptation and the Horse. Cambridge: McDonald Institute. pp. 83–104.Google Scholar
Olsen, S.A. (2006a) Early horse domestication on the Eurasian Steppe. In: Zeder, M.A., Bradley, D.G., Emschwiller, E. and Smith, B.D. (eds), Documenting Domestication: New Genetic and Archaeological Paradigms, Berkeley: University of California Press, pp. 245–69.Google Scholar
Olsen, S.A. (2006b) Early horse domestication: weighing the evidence. In: Olsen, S.L., Grant, S., Choyke, A.M. and Bartosiewicz, L. (eds), Horses and Humans: The Evolution of Human–Equine Relationships. Oxford: Archaeopress, pp. 81113.Google Scholar
Olsen, S.A., Bradley, B., Maki, D. and Outram, A. (2006) Community organization among Copper Age sedentary horse pastoralists of Kazakhstan. In: Peterson, D.L., Popova, L.M. and Smith, A.T. (eds), Beyond the Steppe and the Sown: Proceedings of the 2002 University of Chicago Conference on Eurasian Archaeology. Leiden: Brill. pp. 89111.Google Scholar
Orlando, L. (2015) The first aurochs genome reveals the breeding history of British and European cattle. Genome Biology 16(1), 225.Google Scholar
Orlando, L., Ginolhac, A., Zhang, G., Froese, D., Albrechtsen, A., Stiller, M., et al. (2013) Recalibrating Equus evolution using the genome sequence of an early Middle Pleistocene horse. Nature 499(7456), 74–8.Google Scholar
Orlando, L. and Willerslev, E. (2014) An epigenetic window into the past? Science 345(6196), 511–12.Google Scholar
Ottoni, C., Girdland Flink, L., Evin, A., Geörg, C., De Cupere, B., Van Neer, W., et al. (2013) Pig domestication and human-mediated dispersal in western Eurasia revealed through ancient DNA and geometric morphometrics. Molecular Biology and Evolution 30(4), 824–32.Google Scholar
Outram, A.K. (1998) The identification and Palaeoeconomic context of prehistoric bone marrow and grease exploitation. Unpublished PhD Thesis. University of Durham, England.Google Scholar
Outram, A.K. (1999) A comparison of Paleo-Eskimo and medieval Norse bone fat exploitation in western Greenland. Arctic Anthropology 103–17.Google Scholar
Outram, A.K. (2001) A new approach to identifying bone marrow and grease exploitation: why the indeterminate fragments should not be ignored. Journal of Archaeological Science 28(4), 401–10.Google Scholar
Outram, A.K. (2002) Bone fracture and within-bone nutrients: an experimentally based method for investigating levels of marrow extraction. In: Miracle, P. and Milner, N. (eds), Consuming Passions and Patterns of Consumption, Cambridge: McDonald Institute for Archaeological Research, pp. 5164.Google Scholar
Outram, A.K. (2003) Comparing levels of subsistence stress amongst Norse settlers in Iceland and Greenland using levels of bone fat exploitation as an indicator. Environmental Archaeology 8(2), 119–28.Google Scholar
Outram, A.K. (2004a) Applied models and indices vs. high-resolution, observed data: detailed fracture and fragmentation analyses for the investigation of skeletal part abundance patterns. Journal of Taphonomy 2(3), 167–84.Google Scholar
Outram, A.K. (2004b) Identifying dietary stress in marginal environments: bone fats, optimal foraging theory and the seasonal round. In: Mondini, M., Munoz, S. and Wickler, S. (eds), Colonisation, Migration and Marginal Areas: A Zooarchaeological Approach. Oxford: Oxbow, pp. 7485.Google Scholar
Outram, A.K. (2006) Juggling with indices: a review of the evidence and interpretations regarding upper Palaeolithic horse skeletal part abundances. In: Olsen, S.L., Grant, S., Choyke, A.M. and Bartosiewicz, L. (eds), Horses and Humans: The Evolution of Human–Equine Relationships. Oxford: Archaeopress, pp. 4960.Google Scholar
Outram, A.K. (2014) Animal domestications. In: Cumming, V., Jordan, P. and Zvelebil, M. (eds), Oxford Handbook of the Archaeology and Anthropology of Hunter-Gatherers. Oxford: Oxford University Press, pp. 749–63.Google Scholar
Outram, A.K. (2015) Pastoralism. In: Barker, G. and Goucher, C. (eds), The Cambridge World History, Volume II: A World with Agriculture, 12,000 bce – 500 ce. Cambridge: Cambridge University Press, pp. 161–85.Google Scholar
Outram, A.K. (2017) Answering zooarchaeological questions from the analysis of animal bones and organic pottery residues: a critical comparison. In: Rowley-Conwy, P.A., Serjeantson, D. and Halstead, P. (eds), Economic Zooarchaeology Studies in Hunting, Herding and Early Agriculture. Oxford: Oxbow Books, pp. 148–56.Google Scholar
Outram, A.K., Kasparov, A., Stear, N.A., Varfolomeev, V., Usmanova, E. and Evershed, R.P. (2012) Patterns of pastoralism in later Bronze Age Kazakhstan: new evidence from faunal and lipid residue analyses. Journal of Archaeological Science 39, 2424–35.Google Scholar
Outram, A. and Rowley-Conwy, P. (1998) Meat and marrow utility indices for horse (Equus). Journal of Archaeological Science 25, 839–49.Google Scholar
Outram, A.K., Stear, N.A., Bendrey, R., Olsen, S., Kasparov, A., Zaibert, V., et al. (2009) The earliest horse harnessing and milking. Science 323(5919), 1332–5.Google Scholar
Outram, A.K., Stear, N.A., Kasparov, A., Usmanova, E., Varfolomeev, V. and Evershed, R.P. (2011) Horses for the dead: funerary foodways in Bronze Age Kazakhstan. Antiquity 85, 116–28.Google Scholar
Overton, N.J. and Hamilakis, Y. (2013) A manifesto for a social zooarchaeology. Swans and other beings in the Mesolithic. Archaeological Dialogues 20(2), 111–36.Google Scholar
Pääbo, S. (1985) Molecular cloning of ancient Egyptian mummy DNA. Nature 314, 644–5.Google Scholar
Pääbo, S. (1989) Ancient DNA: extraction, characterization, molecular cloning, and enzymatic amplification. Proceedings of the National Academy of Sciences 86(6), 1939–43.Google Scholar
Palmer, C. and Van der Veen, M. (2002) Archaeobotany and the social context of food. Acta Palaeobotanica 42(2), 195202.Google Scholar
Paludan-Müller, C. (1978) High Atlantic food gathering in northwestern Zealand, ecological conditions and spatial representation. In: Kristiansen, K. and Paludan-Müller, C. (eds), New Directions in Scandinavian Archaeology. Copenhagen: National Museum of Denmark, pp. 120–57.Google Scholar
Park, S.D., Magee, D.A., McGettigan, P.A., Teasdale, M.D., Edwards, C.J., Lohan, A.J., et al. (2015) Genome sequencing of the extinct Eurasian wild aurochs, Bos primigenius, illuminates the phylogeography and evolution of cattle. Genome Biology 16(1), 234.Google Scholar
Parmenter, P.C.R. (2015) A reassessment of the role of animals at the Etton Causewayed Enclosure. PhD Thesis. University of Exeter, England. http://hdl.handle.net/10871/18013.Google Scholar
Parmenter, P.C.R., Johnson, E.V. and Outram, A.K. (2015) Inventing the Neolithic? Putting evidence-based interpretation back into the study of faunal remains from causewayed enclosures. World Archaeology 57(5), 819–33.Google Scholar
Payne, S. (1969) A metrical distinction between sheep and goat metacarpals. In: Ucko, P.J. and Dimbleby, G.W. (eds), The Domestication and Exploitation of Plants and Animals. London: Duckworth, pp. 295305.Google Scholar
Payne, S. (1972a) Partial recover and sample bias: the results of some sieving experiments. In: Higgs, E.S. (ed.), Papers in Economic Prehistory: Studies by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press, pp. 4964.Google Scholar
Payne, S. (1972b) On the interpretation of bone samples from archaeological sites. In: Higgs, E.S. (ed.), Papers in Economic Prehistory: Studies by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press, pp. 2736.Google Scholar
Payne, S. (1973) Kill off patterns in sheep and goats: the mandibles from Asvan Kale. Anatolian Studies 23, 281303.Google Scholar
Pearsall, D.M. (2000) Palaeoethnobotany: A Handbook of Procedures, 2nd edn. San Diego, ca: Academic Press.Google Scholar
Pearsall, D.M., Chandler-Ezell, K. and Chandler-Ezell, A. (2003) Identifying maize in Neotropical sediments and soils using cob phytoliths. Journal of Archaeological Science 30, 611–27.Google Scholar
Pedersen, M.W., Overballe-Petersen, S., Ermini, L., Der Sarkissian, C., Haile, J., Hellstrom, M., et al. (2015) Ancient and modern environmental DNA. Philosphical Transactions of the Royal Society B 370(1660), 20130383.Google Scholar
Pedersen, J.S., Valen, E., Velazquez, A.M.V., Parker, B.J., Rasmussen, M., Lindgreen, S., et al. (2014) Genome-wide nucleosome map and cytosine methylation levels of an ancient human genome. Genome Research 24(3), 454–66.Google Scholar
Peleg, Z., Fahima, T., Korol, A.B., Abbo, S. and Saranga, Y. (2011) Genetic analysis of wheat domestication and evolution under domestication. Journal of Experimental Botany 62(14), 5051–61.Google Scholar
Peña-Chocarro, L. (2007) Early agriculture in central and southern Spain. In: Colledge, S. and Conolly, J. (eds), The origins of domestic plants in Southwest Asia and Europe. Walnut Creek, CA: Left Coast Press, pp. 173–87.Google Scholar
Peng, J.H., Sun, D. and Nevo, E. (2011) Domestication evolution, genetics and genomics in wheat. Molecular Breeding 28(3), 281.Google Scholar
Peukert, S., Bol, R., Roberts, W., Macleod, C.J.A., Murray, P.J., Dixon, E.R. and Brazier, R.E. (2012). Understanding spatial variability of soil properties: a key step in establishing field- to farm-scale agro-ecosystem experiments. Rapid Communications in Mass Spectrometry 26, 2413–21.Google Scholar
Piggott, S. (1992) Wagon, Chariot and Carriage. London: Thames & Hudson.Google Scholar
Piperno, D. (2006) Phytoliths. A Comprehensive Guide for Archaeologists and Palaeoecologists. Lanham, md, New York, Toronto and Oxford: Altamira.Google Scholar
Piperno, D.R. (2017) Assessing elements of an extended evolutionary synthesis for plant domestication and agricultural origin research. Proceedings of the National Academy of Sciences 114(25), 6429–37.Google Scholar
Piperno, D.R. and Holst, I. (1998) The presence of starch grains on prehistoric stone tools from the humid neotropics: indications of early tuber use and agriculture in Panama. Journal of Archaeological Science 25(8), 765–76.Google Scholar
Piperno, D.R., Ranere, A.J., Holst, I., Iriarte, J. and Dickau, R. (2009) Starch grain and phytolith evidence for early ninth millennium B.P. maize from the Central Balsas River valley, Mexico. Proceedings of the National Academy of Sciences 106, 5019–24.Google Scholar
Pitts, M. (1979) Hides and antlers: a new look at the hunter-gatherer site of Star Carr, North Yorksire, England. World Archaeology 11, 3242.Google Scholar
Popper, K. (1959) The Logic of Scientific Discovery. London: Hutchinson.Google Scholar
Privat, K., O’Connell, T., Neal, K. and Hedges, R. (2005) Fermented dairy product analysis and palaeodietary repercussions: is stable isotope analysis not cheesy enough? In: Mulville, J. and Outram, A.K. (eds), The Zooarchaeology of Fats, Oils, Milk and Dairying. Oxford: Oxbow, pp. 60–6.Google Scholar
Quinlan, R.J. (2007) Human parental effort and environmental risk. Proceedings of the Royal Society of London B: Biological Sciences 274(1606), 121–5.Google Scholar
Quintelier, K., Ervynck, A., Müldner, G., Neer, W., Richards, M.P. and Fuller, B.T. (2014) Isotopic examination of links between diet, social differentiation, and DISH at the post-medieval Carmelite Friary of Aalst, Belgium. American Journal of Physical Anthropology 153(2), 203–13.Google Scholar
Ramos-Madrigal, J., Smith, B.D., Moreno-Mayar, J.V., Gopalakrishnan, S., Ross-Ibarra, J., Gilbert, M.T.P. and Wales, N. (2016) Genome sequence of a 5,310-year-old maize cob provides insights into the early stages of maize domestication. Current Biology 26(23), 3195–201.Google Scholar
Ranere, A.J., Piperno, D.R., Holst, I., Dickau, R. and Iriarte, J. (2009) The cultural and chronological context of early Holocene maize and squash domestication in the Central Balsas River Valley. Proceedings of the National Academy of Sciences 106, 5014–18.Google Scholar
Rasmussen, M., Anzick, S.L., Waters, M.R., Skoglund, P., DeGiorgio, M., StaffordJr, T.W., et al. (2014) The genome of a Late Pleistocene human from a Clovis burial site in western Montana. Nature 506(7487), 225–9.Google Scholar
Rasmussen, M., Li, Y., Lindgreen, S., Pedersen, J.S., Albrechtsen, A., Moltke, I., et al. (2010) Ancient human genome sequence of an extinct Palaeo-Eskimo. Nature 463(7282), 757–62.Google Scholar
Reber, E., Baumann, T.E., Monaghan, G. W. and Myers, K.N. (2015) Absorbed Residue Analysis of a Mississippi Plain Jar from Angel Mounds (12Vg1) Lipid Distribution Revisited. Advances in Archaeological Practice 3(1), 2949.Google Scholar
Reber, E.A., Dudd, S.N., Van der Merwe, N.J. and Evershed, R.P. (2004) Direct detection of maize in pottery residues via compound specific stable carbon isotope analysis. Antiquity 78(301), 682–91.Google Scholar
Reber, E. A. and Evershed, R. P. (2004a) How Did Mississippians Prepare Maize? The Application of Compound-Specific Carbon Isotope Analysis to Absorbed Pottery Residues From Several Mississippi Valley Sites. Archaeometry 46(1), 1933.Google Scholar
Reber, E. A. and Evershed, R.P. (2004b) Identification of maize in absorbed organic residues: a cautionary tale. Journal of Archaeological Science 31(4), 399410.Google Scholar
Reidhead, V. (1979) Linear programming models in archaeology. Annual Review of Anthropology 8, 543–78.Google Scholar
Reitmaier, T., Doppler, T., Pike, A.W., Deschler-Erb, S., Hajdas, I., Walser, C. and Gerling, C. (2018) Alpine cattle management during the bronze age at Ramosch-Mottata, Switzerland. Quaternary International 484, 1931.Google Scholar
Reitsema, L.J. (2013) Beyond diet reconstruction: stable isotope applications to human physiology, health, and nutrition. American Journal of Human Biology 25(4), 445–56.Google Scholar
Reitz, E.J. and Wing, E.A. (2008) Zooarchaeology, 2nd ed. Cambridge: Cambridge University Press.Google Scholar
Renfrew, C. (1982) Polity and power: interaction, intensification and exploitation. In: Renfrew, C. and Wagstaff, M. (eds), An Island Polity: the Archaeology of Exploitation on Melos. Cambridge: Cambridge University Press, pp. 264–90.Google Scholar
Renfrew, C. (1987) Archaeology and Language: The Puzzle of Indo-European Origins. London: Jonathan Cape.Google Scholar
Renfrew, C. (2002a) The emerging synthesis: the archaeogenetics of farming/language dispersals and other spread zones. In: Bellwood, P. and Renfrew, C. (eds), Examining the Farming/Language Dispersal Hypothesis. Cambridge: McDonald Institute, pp. 316.Google Scholar
Renfrew, C. (2002b) Pastoralism and interaction: Some introductory questions. In: Boyle, K., Renfrew, C. and Levine, M. (eds), Ancient interactions: East and west in Eurasia. Cambridge: McDonald Institute, pp. 110.Google Scholar
Renfrew, C. and Bahn, B. (2012) Archaeology: Theories, Methods and Practice. 6th Edition. London: Thames & Hudson.Google Scholar
Renfrew, J. (ed.) (1991) New Light on Early Farming: Recent Developments in Palaeoethnobotany. Edinburgh: Edinburgh University Press.Google Scholar
Reuther, J.D., Lowenstein, J.M., Gerlach, S.C., Hood, D., Scheuenstuhl, G. and Ubelaker, D.H. (2006) The use of an improved pRIA technique in the identification of protein residues. Journal of Archaeological Science 33(4), 531–7.Google Scholar
Reynolds, P. (1978) Iron Age Farm: The Butser Experiment. London: British Museum.Google Scholar
Reynolds, P. (1981) Deadstock and Livestock. In: Mercer, R. (ed.), Farming Practice in British Prehistory. Edinburgh: Edinburgh University Press, pp. 97122.Google Scholar
Richards, M.P. and Hedges, R.E.M. (1999) A Neolithic revolution? New evidence of diet in the British Neolithic. Antiquity 73(282), 891–7.Google Scholar
Richards, M.P. and Montgomery, J. (2012) Isotope analysis and palaeopathology: a short review and future developments. In: Buikstra, J. and Roberts, C. (eds), The Global History of Paleopathology: Pioneers and Prospects. Oxford: OUP, pp. 718–31.Google Scholar
Richards, M.P. and Schulting, R.J. (2006) Touch not the fish: the Mesolithic-Neolithic change of diet and its significance. Antiquity 80(308), 444–56.Google Scholar
Richards, T.W. and Lembert, M.E. (1914) The atomic weight of lead of radioactive origin. Journal of the American Chemical Society 36(7), 1329–44.Google Scholar
Riehl, S., Bryson, R. and Pustovoytov, K. (2008) Changing growing conditions for crops during the Near Eastern Bronze Age (3000–1200 bc): the stable carbon isotope evidence. Journal of Archaeological Science 35(4), 1011–22.Google Scholar
Ringrose, T.J. (1993) Bone counts and statistics: a critique. Journal of Archaeological Science 20, 121–57.Google Scholar
Rival, A., Beulé, T., Bertossi, F.A., Tregear, J. and Jaligot, E. (2010) Plant epigenetics: from genomes to epigenomes. Notulae Botanicae Horti Agrobotanici Cluj-Napoca 38(2), 915.Google Scholar
Robb, J. and Miracle, P. (2007) Beyond ‘migration’ versus ‘acculturation’: new models for the spread of agriculture. In: Whittle, and Cummings, V. (eds), Going Over: the Mesolithic-Neolithic Transition in North-west Europe. London: British Academy, pp. 99115.Google Scholar
Robinson, N., Evershed, R.P., Higgs, W.J., Jerman, K. and Eglinton, G. (1987) Proof of a pine wood origin for pitch from Tudor (Mary Rose) and Etruscan shipwrecks: application of analytical organic chemistry in archaeology. Analyst 112(5), 637–44.Google Scholar
Roffet-Salque, M., Lee, M.R., Timpson, A. and Evershed, R.P. (2017) Impact of modern cattle feeding practices on milk fatty acid stable carbon isotope compositions emphasise the need for caution in selecting reference animal tissues and products for archaeological investigations. Archaeological and Anthropological Sciences 9(7), 1343–8.Google Scholar
Roffet-Salque, M., Marciniak, A., Valdes, P.J., Pawłowska, K., Pyzel, J., Czerniak, L., Krüger, M., Roberts, C.N., Pitter, S. and Evershed, R.P. (2018) Evidence for the impact of the 8.2-kyBP climate event on Near Eastern early farmers. Proceedings of the National Academy of Sciences 115(35), 8705–9.Google Scholar
Roffet-Salque, M., Regert, M., Evershed, R.P., Outram, A.K., Cramp, L.J., Decavallas, O., et al. (2015) Widespread exploitation of the honeybee by early Neolithic farmers. Nature 527(7577), 226–30.Google Scholar
Roper, D.C. (1979) The method and theory of site catchment analysis: a review. Advances in Archaeological Method and Theory 2, 119–40.Google Scholar
Rösch, M., Kleinmann, A., Lechterbeck, J. and Wick, L. (2014) Botanical off-site and on-site data as indicators of different land use systems: A discussion with examples from Southwest Germany. Vegetation History and Archaeobotany 23, 121–33.Google Scholar
Rosen, A. and Weiner, S. (1994) Identifying ancient irrigation: a new method using opaline phytoliths from emmer wheat. Journal of Archaeological Science 21, 125–32.Google Scholar
Rossman, D.L. (1976) A site catchment analysis of San Lorenzo, Veracruz. In: Flannery, K.V. (ed.), The Early Mesoamerican Village. New York: Academic Press, pp. 95103.Google Scholar
Rottoli, M. and Pessina, A. (2007) Early agriculture in central and southern Spain. In: Colledge, S. and Conolly, J. (eds), The Origins and Spread of Domestic Plants in Southwest Asia and Europe. Walnut Creek, ca: Left Coast Press, pp. 141–53.Google Scholar
Rowley-Conwy, P.A. (1981a) Mesolithic Danish bacon: permanent and temporary sites in the Danish Mesolithic. In: Bailey, G. and Sheridan, A. (eds), Economic Archaeology. Oxford: British Archaeological Reports, pp. 51–8.Google Scholar
Rowley-Conwy, P.A. (1981b) Slash and burn in the temperate European Neolithic. In: Mercer, R. (ed.), Farming Practice in British Prehistory. Edinburgh: Edinburgh University Press, pp. 8596.Google Scholar
Rowley-Conwy, P.A. (1983) Sedentary hunters: the Ertebølle example. In: Bailey, G. (ed.), Hunter-Gatherer Economy in Prehistory: A European Perspective. Cambridge: Cambridge University Press, pp. 111–26.Google Scholar
Rowley-Conwy, P.A. (1986) Between cave painters and crop planters: aspects of the temperate European Mesolithic. In: Zvelebil, M. (ed.), Hunters in Transition: Mesolithic Societies of Temperate Eurasia and their Transition to Farming. Cambridge: Cambridge University Press, pp. 1732.Google Scholar
Rowley-Conwy, P.A. (1987) Animal bones in mesolithic studies: recent progress and hopes for the future. In: Rowley-Conwy, P., Zvelebil, M. and Blankholm, H.P. (eds.), Mesolithic Northwest Europe – Recent Trends. Sheffield: Department of Archaeology and Prehistory, pp. 7481.Google Scholar
Rowley-Conwy, P.A. (1994) Dung, dirt and deposits: site formation under conditions of near-perfect preservation at Qasr Ibrim, Egyptian Nubia. In: Luff, R. and Rowley-Conwy, P.A. (eds), Whither Environmental Archaeology? Oxford: Oxbow, pp. 2532.Google Scholar
Rowley-Conwy, P.A. (1998a) Meat, furs and skins: Mesolithic animal bones from Ringkloster, a seasonal hunting camp in Jutland. Journal of Danish Archaeology v.12 1994–95, 8798.Google Scholar
Rowley-Conwy, P.A. (1998b) Cemeteries, seasonality and complexity in the Ertebølle of Southern Scandinavia. In: Zvelebil, M., Dennell, R. and Domańska, L. (eds), Harvesting the Sea, Farming the Forest: The Emergence of Neolithic Societies in the Baltic region. Sheffield: Sheffield Academic Press, pp. 193202.Google Scholar
Rowley-Conwy, P.A. (1999) Economic prehistory in Southern Scandinavia. In: Coles, J., Bewley, R. and Mellars, P. (eds), World Prehistory: Studies in Memory of Grahame Clark. Oxford: Oxford University Press, pp. 125–59.Google Scholar
Rowley-Conwy, P.A. (2001) Time, change and the archaeology of hunter-gatherers: how original is the ‘Original Affluent Society’? In: Panter-Brick, C., Layton, R.H. and Rowley-Conwy, P. (eds), Hunter-Gatherers: An Interdisciplinary Perspective. Cambridge: Cambridge University Press, pp. 3972.Google Scholar
Rowley-Conwy, P., Halstead, P. and Collins, P. (2002) Derivation and application of a food utility index (FUI) for European wild boar (Sus scrofa L.). Environmental Archaeology 7(1), 7788.Google Scholar
Rowley-Conwy, P. and Layton, R. (2011) Foraging and farming as niche construction: stable and unstable adaptations. Philosophical Transactions of the Royal Society B: Biological Sciences 366(1566), 849–62.Google Scholar
Rowley-Conwy, P. and Storå, J. (1997) Pitted Ware seals and pigs from Ajvide, Gotland: methods of study and first results. In: Burenhult, G. (ed.), Remote Sensing 1, Stockholm: Stockholm University, pp. 113–25.Google Scholar
Russell, N. (2012) Social Zooarchaeology: Humans and Animals in Prehistory. Cambridge: Cambridge University Press.Google Scholar
Rutgers, L.V., Van Strydonck, M., Boudin, M. and Van der Linde, C. (2009) Stable isotope data from the early Christian catacombs of ancient Rome: new insights into the dietary habits of Rome’s early Christians. Journal of Archaeological Science 36(5), 1127–34.Google Scholar
Saag, L., Varul, L., Scheib, C.L., Stenderup, J., Allentoft, M.E., Saag, L., Pagani, L., Reidla, M., Tambets, K., Metspalu, E., Kriiska, A., Willerslev, E., Kivisild, T. and Metspalu, M. (2017) Extensive farming in Estonia started through a sex-biased migration from the Steppe. Current Biology 27(14), 2185–93.Google Scholar
Sahlins, M. (1974) Stone Age Economics. London: Tavistock.Google Scholar
Salque, M., Bogucki, P.I., Pyzel, J., Sobkowiak-Tabaka, I., Grygiel, R., Szmyt, M. and Evershed, R.P. (2013). Earliest evidence for cheese making in the sixth millennium bc in northern Europe. Nature 493(7433), 522–5.Google Scholar
Samuel, D. (2006) Modified starch. In: Torrence, R. and Barton, H. (eds), Ancient Starch Research. Walnut Creek, ca: Left Coast Press, pp. 205–16.Google Scholar
Sanger, F. and Coulson, A.R. (1975). A rapid method for determining sequences in DNA by primed synthesis with DNA polymerase. Journal of Molecular Biology 94(3), 441IN19447–446IN20448.Google Scholar
Sanger, F., Nicklen, S. and Coulson, A.R. (1977) DNA sequencing with chain-terminating inhibitors. Proceedings of the National Academy of Sciences 74(12), 5463–7.Google Scholar
Sano, K. (2012) Functional variability in the Magdalenian of north-western Europe: A lithic microwear analysis of the Gönnersdorf K-II assemblage. Quaternary International 272, 264–74.Google Scholar
Savelle, J.M. and Friesen, M.T. (1996) An Odontocete (Cetacea) meat utility index. Journal of Archaeological Science 23(5), 713–21.Google Scholar
Schäfer, M. (2010). Archäozoologische Untersuchung der Tierknochen aus der linearbandkeramischen Siedlung Vaihingen/Enz (Kreis Ludwigsburg D) und ihre Interpretation. Unpublished PhD, Basel University.Google Scholar
Schaefer, N.K., Shapiro, B. and Green, R.E. (2016) Detecting hybridization using ancient DNA. Molecular Ecology 25(11), 2398–412.Google Scholar
Schibler, J. (2001). Methodische Überlegungen zum Problem der Einschätzung der Bedeutung von Jagd und Viehwirtschaft im schweizerischen Neolithikum. In: Arbogast, R.-M., Jeunesse, C. and Schibler, J. (eds), Premières rencontres danubiennes, Strasbourg 20 et 21 novembre 1996, Actes de la première table-ronde; Rôle et statut de la chasse dans le Néolithique ancien danubien (5500–4900 av. J.-C.). Rahden: Verlag Marie Leidorf, pp. 153–70.Google Scholar
Schibler, J., Hüster-Plogmann, H., Jacomet, S., Brombacher, C., Gross-Klee, E. and Rast-Eicher, A. (1997) Ökonomie und Ökologie neolithischer und bronzezeitlicher Ufersiedlungen am Zürichsee. Zurich: Zürich und Egg.Google Scholar
Schier, W. (2009) Extensiver Brandfeldbau und die Ausbreitung der neolithischen Wirtschaftsweise in Mitteleuropa und Südskandinavien am Ende der 5. Jahrtausends v. Chr. Praehistorische Zeitschrift 84, 1543.Google Scholar
Schmaus, T.M., Chang, C. and Tourtellotte, P.A. (2018) A model for pastoral mobility in Iron Age Kazakhstan. Journal of Archaeological Science: Reports 17, 137–43.Google Scholar
Schmid, E. (1972) Atlas of Animal Bones: For Prehistorians, Archaeologists, and Quaternary Geologists. Amsterdam: Elsevier.Google Scholar
Schoeninger, M.J. and DeNiro, M.J. (1984) Nitrogen and carbon isotopic composition of bone collagen from marine and terrestrial animals. Geochimica et Cosmochimica Acta 48(4), 625–39.Google Scholar
Schoville, B.J. and Otárola-Castillo, E. (2014) A model of hunter-gatherer skeletal element transport: the effect of prey body size, carriers, and distance. Journal of Human Evolution 73, 114.Google Scholar
Schurr, M.R. (1998) Using stable nitrogen-isotopes to study weaning behavior in past populations. World Archaeology 30(2), 327–42.Google Scholar
Schutkowski, H. (2006) Human Ecology: Biocultural Adaptations in Human Communities. Heidelberg: Springer.Google Scholar
Sealy, J. (2001) Body tissue chemistry and palaeodiet. In: Brothwell, D.R. and Pollard, A.M. (eds), Handbook of Archaeological Sciences. Chichester: Wiley, pp. 269–79.Google Scholar
Seetah, K. (2008) Modern analogy, cultural theory and experimental replication: a merging point at the cutting edge of archaeology. World Archaeology 40(1), 135–50.Google Scholar
Seetah, K., Cucchi, T., Dobney, K. and Barker, G. (2014) A geometric morphometric re-evaluation of the use of dental form to explore population differences in horses (Equus caballus) and its potential zooarchaeological application. Journal of Archaeological Science 41, 904–10.Google Scholar
Ségurel, L. and Bon, C. (2017) On the Evolution of Lactase Persistence in Humans. Annual Review of Genomics and Human Genetics 18(8.1), 8.23Google Scholar
Seguin-Orlando, A., Gamba, C., Der Sarkissian, C., Ermini, L., Louvel, G., Boulygina, E., et al. (2015) Pros and cons of methylation-based enrichment methods for ancient DNA. Scientific Reports 5, 11826.Google Scholar
Seinfeld, D.M., von Nagy, C. and Pohl, M.D. (2009) Determining Olmec maize use through bulk stable carbon isotope analysis. Journal of Archaeological Science 36(11), 2560–5.CrossRefGoogle Scholar
Service, E.R. (1962) Primitive Social Organization: An Evolutionary Perspective. New York: Random House.Google Scholar
Service, E.R. (1966) The Hunters. Englewood Cliffs: Prentice-Hall.Google Scholar
Sharp, Z.D., Atudorei, V., Panarello, H.O., Fernández, J. and Douthitt, C. (2003) Hydrogen isotope systematics of hair: archeological and forensic applications. Journal of Archaeological Science 30(12), 1709–16.CrossRefGoogle Scholar
Shaw, H., Montgomery, J., Redfern, R., Gowland, R. and Evans, J. (2016) Identifying migrants in Roman London using lead and strontium stable isotopes. Journal of Archaeological Science 66, 5768.Google Scholar
Sherratt, A. (1980). Water, soil and seasonality in early cereal cultivation. World Archaeology 2, 313–30.Google Scholar
Sherratt, A.G. (1981) Plough and pastoralism, aspects of the secondary products revolution, In: Hodder, I., Isaac, G. and Hammond, N. (eds.), Patterns of the Past. Cambridge: Cambridge University Press, pp. 261306.Google Scholar
Sherratt, A.G. (1983) The secondary products revolution of animals in the Old World, World Archaeology 15, 90104.Google Scholar
Sherratt, A. (2006) The Trans-Eurasian exchange: the prehistory of Chinese relations with the West. In: Mair, V. (ed.), Contact and Exchange in the Ancient World. Honolulu: Hawaii University Press, pp. 3253.Google Scholar
Shillito, L.M. and Ryan, P. (2013) Surfaces and streets: Phytoliths, micromorphology and changing use of space at neolithic Çatalhöyük (Turkey). Antiquity 87(337), 684700.Google Scholar
Shishlina, N. (2003) Yamnaya culture pastoral exploitation: a local sequence. In: Levine, M., Renfrew, C. and Boyle, K. (eds), Prehistoric Steppe Adaptation and the Horse. Cambridge: McDonald Institute, pp. 353–65.Google Scholar
Sielmann, B. (1971) Der Einfluß der Umwelt auf die neolithische Besiedlung Südwestdeutschlands unter besonderer Berücksichtigung der Verhältnisse am nördlichen Oberrhein. Acta praehistorica et Archaeologica 2, 65197.Google Scholar
Silver, I.A. (1969) The ageing of domestic animals. In: Brothwell, D. and Higgs, E. (eds), Science in Archaeology. London: Thames & Hudson, pp. 283302.Google Scholar
Sluyter, A. (2003) Neo-Environmental Determinism, Intellectual Damage Control, and Nature/Society Science. Antipode 35(4), 813–17.Google Scholar
Smith, B.D. (2007) Niche construction and the behavioral context of plant and animal domestication. Evolutionary Anthropology: Issues, News, and Reviews 16(5), 188–99.Google Scholar
Smith, B.D. (2009) Core conceptual flaws in human behavioral ecology. Communicative and Integrative Biology 2(6), 533–4.Google Scholar
Smith, B.D. (2011) A cultural Niche Construction Theory of initial domestication. Biological Theory 6(3), 260–71.Google Scholar
Smith, B.D. and Zeder, M.A. (2013) The onset of the anthropocene. Anthropocene 4, 813.Google Scholar
Smith, E.A. (1983) Comment on: Territoriality among human foragers: ecological models and an application to four Bushman groups. Current Anthropology 24(1), 61.Google Scholar
Smith, P.R. and Wilson, M.T. (2001) Blood residues in archaeology. In: Brothwell, D.R. and Pollard, A.M. (eds), Handbook of Archaeological Sciences. Chichester: Wiley, pp. 313–22.Google Scholar
Sobolik, K.D. and Steele, D.G. (1996) A Turtle Atlas to Facilitate Archaeological Identifications. Rapid City: Mammoth Site of Hot Springs and Kristin Sobolik.Google Scholar
Soddy, F. (1913) Intra-atomic charge. Nature 92(2301), 399400.Google Scholar
Spangenberg, J.E., Matuschik, I., Jacomet, S. and Schibler, J. (2008) Direct evidence for the existence of dairying farms in prehistoric Central Europe (4th millennium bc). Isotopes in Environmental and Health Studies 44, 189200.Google Scholar
Speth, J.D. (1983) Bison Kills and Bone Counts: Decision Making by Ancient Hunters. Chicago: University of Chicago Press.Google Scholar
Speth, J.D. and Spielmann, K.A. (1983) Energy source, protein metabolism, and hunter-gatherer subsistence strategies. Journal of Anthropological Archaeology 2, 131.Google Scholar
Spigelman, M., Donoghue, H.D., Abdeen, Z., Ereqat, S., Sarie, I., Greenblatt, C.L., et al. (2015) Evolutionary changes in the genome of Mycobacterium tuberculosis and the human genome from 9000 years bp until modern times. Tuberculosis 95, S145S149.Google Scholar
Stear, N.A. (2008) Changing patterns of animal exploitation in the prehistoric Eurasian steppe: an integrated molecular, stable isotope and archaeological approach, Unpublished Thesis, University of Bristol.Google Scholar
Stewart, N.A., Gerlach, R.F., Gowland, R.L., Gron, K.J. and Montgomery, J. (2017) Sex determination of human remains from peptides in tooth enamel. Proceedings of the National Academy of Sciences 114(52), 13649–54.Google Scholar
Stiner, M.C. (2001) Thirty years on the ‘Broad Spectrum Revolution’ and Palaeolithic demography. Proceedings of the National Academy of Sciences 98(13), 6993–6.Google Scholar
Stott, A.W., Berstan, R., Evershed, R.P., Bronk-Ramsey, C., Hedges, R.E. and Humm, M.J. (2003) Direct dating of archaeological pottery by compound-specific 14C analysis of preserved lipids. Analytical Chemistry 75(19), 5037–45.Google Scholar
Stott, A., Berstan, R., Evershed, P., Hedges, R., Ramsey, C. B. and Humm, M. (2001) Radiocarbon dating of single compounds isolated from pottery cooking vessel residues. Radiocarbon 43(2A), 191–7.Google Scholar
Street, J. (1969) An evaluation of the concept of carrying capacity. Professional Geographer 21, 104–7.Google Scholar
Strien, H.-C. (2005) Familientraditionen in der bandkeramischen Siedlung bei Vaihingen/Enz. In: Lüning, J., Fridrich, C. and Zimmerman, A. (eds), Die Bandkeramik im 21. Jahrhundert: Symposium in der Abtei Brauweiler bei Köln vom 16.9.-19.9.2002. Rahden: Verlag Marie Leidorf, pp. 189–97.Google Scholar
Strien, H.-C. (2011) Chronological and social interpretation of the artefactual assemblage. In: Bogaard, A., Plant Use and Crop Husbandry in an Early Neolithic Village: Vaihingen an der Enz, Baden-Württemberg. Bonn: Habelt, pp. 1923.Google Scholar
Strien, H.-C. (2014) Eine neue Seriation der ältesten Linienbandkeramik: Zeitliche und räumliche Differenzierung. In: Beier, H.-J., Einicke, R. and Biermann, E. (eds), Material – Werkzeug: Werkzeug – Material & Klinge, Messer, Schwert und Co. – Neues aus der Schneidenwelt. Aktuelles aus der Neolithforschung. Beiträge der Tagungen der Arbeitsgemeinschaft Werkzeuge und Waffen Pottenstein (Fränkische Schweiz) 2011 & Herxheim bei Landau in der Pfalz 2012 sowie Aktuelles. Varia neolithica VIII, pp. 141–61.Google Scholar
Sturdy, D.A. (1975) Some reindeer economies in prehistoric Europe. In: Higgs, E.S. (ed.), Palaeoeconomy: Being the Second Volume of Papers in Economic Prehistory by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press, pp. 5596.Google Scholar
Styring, A.K., Ater, M., Hmimsa, Y., Fraser, R., Miller, H., Neef, R., et al. (2016a). Disentangling the effect of farming practice from aridity on crop stable isotope values: A present-day model from Morocco and its application to early farming sites in the eastern Mediterranean. The Anthropocene Review 3: 222Google Scholar
Styring, A.K., Charles, M., Fantone, F., Hald, M.M., McMahon, A., Meadow, R.H., et al. (2017a) Isotope evidence for agricultural extensification reveals how the world’s first cities were fed. Nature Plants 3(6), 17076.Google Scholar
Styring, A.K., Fraser, R.A., Arbogast, R.M., Halstead, P., Isaakidou, V., Pearson, J.A., et al. (2015) Refining human palaeodietary reconstruction using amino acid δ15N values of plants, animals and humans. Journal of Archaeological Science 53, 504–15.Google Scholar
Styring, A.K., Fraser, R.A., Bogaard, A. and Evershed, R.P. (2014a) The effect of manuring on cereal and pulse amino acid delta N-15 values. Phytochemistry 102, 40–5.Google Scholar
Styring, A.K., Fraser, R.A., Bogaard, A. and Evershed, R.P. (2014b) Cereal grain, rachis and pulse seed amino acid delta N-15 values as indicators of plant nitrogen metabolism. Phytochemistry 97, 20–9.Google Scholar
Styring, A., Maier, U., Stephan, E., Schlichtherle, H. and Bogaard, A. (2016) Cultivation of choice: new insights into farming practices at Neolithic lakeshore sites. Antiquity 90, 95110.Google Scholar
Styring, A., Manning, H., Fraser, R., Wallace, M., Jones, G., Charles, M., et al. (2013) The effect of charring and burial on the biochemical composition of cereal grains: investigating the integrity of archaeological plant material. Journal of Archaeological Science 40, 4767–79.Google Scholar
Styring, A., Rösch, M., Stephan, E., Stika, H.-P., Fischer, E., Sillmann, M. and Bogaard, A. (2017b) Centralisation and long-term change in farming regimes: comparing agricultural practice in Neolithic and Iron Age south-west Germany. Proceedings of the Prehistoric Society 83, 357–81.Google Scholar
Sykes, N. (2004) The introduction of fallow deer to Britain: a zooarchaeological perspective. Environmental Archaeology 9(1), 7583.Google Scholar
Sykes, N. (2014) Beastly Questions: Animal Answers to Archaeological Issues. London: Bloomsbury.Google Scholar
Sykes, N.J., White, J., Hayes, T.E. and Palmer, M.R. (2006) Tracking animals using strontium isotopes in teeth: the role of fallow deer (Dama dama) in Roman Britain. Antiquity 80(310), 948–59.Google Scholar
Taché, K. and Craig, O.E. (2015) Cooperative harvesting of aquatic resources and the beginning of pottery production in north-eastern North America. Antiquity 89(343), 177–90.Google Scholar
Tarasov, P.E., Jolly, D. and Kaplan, J.O. (1997) A continuous Late Glacial and Holocene record of vegetation changes in Kazakhstan. Palaeogeography, Palaeoclimatology, Palaeoecology 136(1–4), 281–92.Google Scholar
Tauber, H. (1981) 13C evidence for dietary habits of prehistoric man in Denmark. Nature 292, 332–3.Google Scholar
Taylor, W.T.T., Bayarsaikhan, J. and Tuvshinjargal, T. (2015) Equine cranial morphology and the identification of riding and chariotry in late Bronze Age Mongolia. Antiquity 89(346), 854–71.CrossRefGoogle Scholar
Thomas, J. (1991) Rethinking the Neolithic. Cambridge: Cambridge University Press.Google Scholar
Thomas, J. (1999) Understanding the Neolithic. London: Routledge.Google Scholar
Thomas, J. (2003) Thoughts on the ‘repacked’ Neolithic revolution. Antiquity 77(295), 6774.Google Scholar
Thomson, J.J. (1913) Bakerian lecture: Rays of positive electricity. Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character 89(607), 120.Google Scholar
Tilley, C. (1981) Economy and society: what relationship? In: Bailey, G. and Sheridan, A. (eds), Economic Archaeology. Oxford: British Archaeological Reports, pp. 131–48.Google Scholar
Tito, R.Y., Knights, D., Metcalf, J., Obregon-Tito, A.J., Cleeland, L., Najar, F., et al. (2012) Insights from characterizing extinct human gut microbiomes. PloS One 7(12), e51146.Google Scholar
Tito, R.Y., Macmil, S., Wiley, G., Najar, F., Cleeland, L., Qu, C., et al. (2008) Phylotyping and functional analysis of two ancient human microbiomes. PLoS One 3(11), e3703.Google Scholar
Torrence, R. (2006) Starch and archaeology. In: Torrence, R. and Barton, H. (eds), Ancient Starch Research, Walnut Creek, ca: Left Coast Press, pp. 1733.Google Scholar
Towers, J., Gledhill, A., Bond, J. and Montgomery, J. (2014) An investigation of cattle birth seasonality using δ13C and δ18O profiles within first molar enamel. Archaeometry 56(S1), 208–36.Google Scholar
Tresset, A. and Vigne, J.-D. (2001) La chasse, principal élément structurant la diversité des faunes archéologiques du Néolithique ancien, en Europe tempérée at en Méditerranée: tentative d’interpretation fonctionelle. In: Arbogast, R.-M., Jeunesse, C. and Schibler, J. (eds), Premières rencontres danubiennes, Strasbourg 20 et 21 novembre 1996, Actes de la première table-ronde; Rôle et statut de la chasse dans le Néolithique ancien danubien (5500–4900 av. J.-C.). Rahden: Verlag Marie Leidorf, pp. 129–51.Google Scholar
Trigger, B.G. (1989) A History of Archaeological Thought. Cambridge: Cambridge University Press.Google Scholar
Troy, C.S., MacHugh, D.E., Bailey, J.F. and Magee, D.A. (2001) Genetic evidence for Near-Eastern origins of European cattle. Nature 410(6832), 1088–91.Google Scholar
Tsalkin, V.I. (1972) Fauna iz raskopok andronovskih pamiatnikov v Priural’e, Osnovnye problemy teriologii. Proceedings of MOIP 48, 6681.Google Scholar
Turner, B.L., Kamenov, G.D., Kingston, J.D. and Armelagos, G.J. (2009) Insights into immigration and social class at Machu Picchu, Peru based on oxygen, strontium, and lead isotopic analysis. Journal of Archaeological Science 36(2), 317–32.Google Scholar
Tuross, N., Barnes, I. and Potts, R. (1996) Protein identification of blood residues on experimental stone tools. Journal of Archaeological Science 23(2), 289–96.Google Scholar
Tykot, R.H. (2004) Stable Isotopes and Diet: You Are What You Eat. In: Martini, M., Milazzo, M. and Piacentini, M. (eds), Physics Methods in Archaeometry. Proceedings of the International School of Physics ‘Enrico Fermi’ Course CLIV, Bologna, Italy: Società Italiana di Fisica, pp. 433–44.Google Scholar
Ugent, D., Pozorski, S. and Pozorski, T. (1981) Prehistoric remains of the sweet potato from the Casma Valley of Peru. Phytologia 49, 401–15.Google Scholar
Usmanova, E (2005) Mogilnik Lisakovskij: Fakty i Parallyeli. Lisakovsk: Lisakovsk Museum.Google Scholar
Vaiglova, P., Snoeck, C., Nitsch, E., Bogaard, A. and Lee-Thorp, J. (2014) Impact of contamination and pre-treatment on stable carbon and nitrogen isotopic composition of charred plant remains. Rapid Communications in Mass Spectrometry 28, 2497–510.Google Scholar
Valamoti, S., Samuel, D., Bayram, M. and Marinova, E. (2008) Prehistoric cereal foods from Greece and Bulgaria: investigation of starch microstructure in experimental and archaeological charred remains. Vegetation History and Archaeobotany 17 (Suppl 1), S265S276.Google Scholar
Vallebueno-Estrada, M., Rodríguez-Arévalo, I., Rougon-Cardoso, A., González, J.M., Cook, A.G., Montiel, R. and Vielle-Calzada, J.-P. (2017) The earliest maize from San Marcos Tehuacan is a partial domesticate with genomic evidence of inbreeding. Proceedings of the National Academy of Sciences 113, 14151–6.Google Scholar
van der Merwe, N.J. and Vogel, J.C. (1978) 13C content of human collagen as a measure of prehistoric diet in woodland North America. Nature 276, 815–16.Google Scholar
van Zeist, W. (1969) Reflections on prehistoric environment in the Near East. In: Ucko, P.J. and Dimbleby, G.W. (eds), The Domestication and Exploitation of Plants and Animals. London: Duckworth, pp. 3546.Google Scholar
van Zeist, W. and Casperie, W.A. (eds) (1984) Plants and Ancient Man: Studies in Palaeoethnobotany. Rotterdam: Balkema.Google Scholar
Ventresca Miller, A.R., Winter-Schuh, C., Usmanova, E.R., Logvin, A., Shevnina, I. and Makarewicz, C.A. (2017) Pastoralist Mobility in Bronze Age Landscapes of Northern Kazakhstan: 87Sr/86Sr and δ18O Analyses of Human Dentition from Bestamak and Lisakovsk. Environmental Archaeology, DOI: 10.1080/14614103.2017.1390031.Google Scholar
Verbricky-Todd, E. (1984) Communal Buffalo Hunting among the Plains Indians. Edmonton: Archaeological Survey of Alberta.Google Scholar
Vigne, J-D. and Helmer, D. (2007) Was milk a ‘secondary product’ in the Old World neolithisation process? Its role in the domestication of cattle, sheep and goats. Anthropozoologica 42, 940.Google Scholar
Vilà, C., Leonard, J.A. and Beja-Pereira, A. (2006) Genetic documentation of horse and donkey domestication. In: Zeder, M.A., Bradley, D.G., Emschwiller, E. and Smith, B.D. (eds), Documenting domestication: new genetic and archaeological paradigms, Berkeley: University of California Press, pp. 342–53.Google Scholar
Vilà, C., Leonard, A., Götherström, S., Marklund, J., Sandberg, K., Lidén, R., et al. (2001) Widespread origins of domestic horse lineages. Science 291, 474–7.Google Scholar
Vita-Finzi, C. and Higgs, E.S. (1970) Prehistoric economy in the Mount Carmel area of Palestine: site catchment analysis. Proceedings on the Prehistoric Society 36, 137.Google Scholar
Vogel, J.C. and Van der Merwe, N.J. (1977) Isotopic evidence for early maize cultivation in New York State. American Antiquity 42, 238–42.Google Scholar
von den Driesch, A.E. (1976) A Guide to the Measurement of Animal Bones from Archaeological Sites. Cambridge, ma: Peabody Museum.Google Scholar
von Holstein, I.C., Rogers, P.W., Craig, O.E., Penkman, K.E., Newton, J. and Collins, M.J. (2016) Provenancing Archaeological Wool Textiles from Medieval Northern Europe by Light Stable Isotope Analysis (δ13C, δ15N, δ2H). PloS one 11(10), p.e0162330.Google Scholar
Wada, E., Mizutani, H. and Minagawa, M. (1991) The use of stable isotopes for food web analysis. Critical Reviews in Food Science & Nutrition 30(4), 361–71.Google Scholar
Wall, J.D. and Kim, S.K. (2007) Inconsistencies in Neanderthal genomic DNA sequences. PLoS Genetics 3(10), p.e175.Google Scholar
Wallace, M., Jones, G., Charles, M., Fraser, R., Halstead, P., Heaton, T.H. and Bogaard, A. (2013) Stable carbon isotope analysis as a direct means of inferring crop water status and water management practices. World Archaeology 45(3), 388409.Google Scholar
Wang, C.C., Farina, S.E. and Li, H. (2013) Neanderthal DNA and modern human origins. Quaternary International 295, 126–9.Google Scholar
Warinner, C., Hendy, J., Speller, C., Cappellini, E., Fischer, R., Trachsel, C., et al. (2014) Direct evidence of milk consumption from ancient human dental calculus. Scientific Reports 4, 7104.Google Scholar
Warinner, C., Speller, C. and Collins, M.J. (2015a) A new era in palaeomicrobiology: prospects for ancient dental calculus as a long-term record of the human oral microbiome. Philosophical Transactions of the Royal Society B 370(1660), 20130376.Google Scholar
Warinner, C., Speller, C., Collins, M.J. and Lewis, C.M. (2015b) Ancient human microbiomes. Journal of Human Evolution 79, 125–36.Google Scholar
Watson, J.D. and Crick, F.H. (1953) Molecular structure of nucleic acids. Nature 171(4356), 737–8.CrossRefGoogle ScholarPubMed
Warmuth, V., Eriksson, A., Bower, M.A., Barker, G., Barrett, E., Hanks, B.K., et al. (2012) Reconstructing the origin and spread of horse domestication in the Eurasian steppe Proceedings from the National Academy of Sciences 109(21), 8202–6.Google Scholar
Webb, E.A. and Longstaffe, F.J. (2000) The oxygen isotopic compositions of silica phytoliths and plant water in grasses: Implications for the study of paleoclimate. Geochemica et Cosmochimica 64, 767–80.Google Scholar
Webb, E.C., Lewis, J., Shain, A., Kastrisianaki-Guyton, E., Honch, H.V., Stewart, A., et al. (2017) The influence of varying proportions of terrestrial and marine dietary protein on the stable carbon-isotope compositions of pig tissues from a controlled feeding experiment. Science and Technology of Archaeological Research 3(1), 3652.Google Scholar
Webb, E.C., Honch, N.V., Dunn, P.J., Linderholm, A., Eriksson, G., Lidén, K. and Evershed, R.P. (2018) Compound-specific amino acid isotopic proxies for distinguishing between terrestrial and aquatic resource consumption. Archaeological and Anthropological Sciences 10(1), 118.Google Scholar
Webb, S.D. and Hemmings, C.A. (2006) Last horses and first humans in North America. In: Olsen, S.L., Grant, S., Choyke, A.M. and Bartosiewicz, L. (eds), Horses and humans: the evolution of human-equine relationships. Oxford: Archaeopress, pp. 1123.Google Scholar
Webley, D. (1972) Soils and site location in prehistoric Palestine. In: Higgs, E.S. (ed.), Papers in Economic Prehistory: Studies by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press, pp. 169–80.Google Scholar
Weisskopf, A., Harvey, E., Kingwell-Banham, E., Kajale, M., Mohanty, R. and Fuller, D.Q. (2014) Archaeobotanical implications of phytolith assemblages from cultivated rice systems, wild rice stands and macro-regional patterns. Journal of Archaeological Science 51, 4353.Google Scholar
Weisskopf, A.R. and Lee, G.A. (2016) Phytolith identification criteria for foxtail and broomcorn millets: a new approach to calculating crop ratios. Archaeological and Anthropological Sciences 8(1), 2942.Google Scholar
Weisskopf, A., Qin, L., Ding, J., Ding, P., Sun, G. and Fuller, D.Q. (2015) Phytoliths and rice: from wet to dry and back again in the Neolithic Lower Yangtze. Antiquity 89, 1051–63.Google Scholar
Wendrich, W. and Barnard, H. (2008) The archaeology of mobility: definitions and research approaches. In: Barnard, H. and Wendrich, W. (eds), The Archaeology of Mobility: Old World and New World Nomadism. Los Angeles: Cotsen Institute, University of California, pp. 124.Google Scholar
West, D. (2006) Horse hunting in Central Europe at the end of the Pleistocene. In: Olsen, S.L., Grant, S., Choyke, A.M. and Bartosiewicz, L. (eds), Horses and Humans: The Evolution of Human-equine Relationships. Oxford: Archaeopress, pp. 2547.Google Scholar
West, J.A. and Louys, J. (2007) Differentiating bamboo from stone tool cut marks in the zooarchaeological record, with a discussion on the use of bamboo knives. Journal of Archaeological Science 34(4), 512–18.Google Scholar
White, T.D. (1992) Prehistoric Cannibalism at Mancos 5MTUMR-2346. Princeton: Princeton University Press.Google Scholar
White, T.E. (1952) Animal bone and plains archaeology. Plains Archaeological Conference Newsletter 4, 46–8.Google Scholar
White, T.E. (1953) Observations on the butchering technique of some aboriginal peoples. American Antiquity 19, 160–4.Google Scholar
Whittle, A. (1996) Europe in the Neolithic. Cambridge: Cambridge University Press.Google Scholar
Whittle, A. (1997) Moving on and moving around: Neolithic settlement mobility. In: Topping, P. (ed.), Neolithic Landscapes. Oxford: Oxbow, pp. 1522.Google Scholar
Widga, C., Walker, J.D. and Stockli, L.D. (2010) Middle Holocene Bison diet and mobility in the eastern Great Plains (USA) based on δ13C, δ18O, and 87Sr/86Sr analyses of tooth enamel carbonate. Quaternary Research 73(3), 449–63.Google Scholar
Wilkinson, P.F. (1972) Current experimental domestication and its relevance to prehistory. In: Higgs, E.S. (ed.), Papers in Economic Prehistory: Studies by Members and Associates of the British Academy Major Research Project in the Early History of Agriculture. Cambridge: Cambridge University Press, pp. 10718.Google Scholar
Willerslev, R., Vitebsky, P. and Alekseyev, A. (2015) Sacrifice as the ideal hunt: a cosmological explanation for the origin of reindeer domestication. Journal of the Royal Anthropological Institute 21(1), 123.Google Scholar
Winterhalder, B. (1986) Diet choice, risk, and food sharing in a stochastic environment. Journal of Anthropological Archaeology 5(4), 369–92.Google Scholar
Woodward, S.R., Weyand, N.J. and Bunnell, M. (1994) DNA sequence from Cretaceous period bone fragments. Science 266(5188), 1229–32.Google Scholar
Wu, Y., Jiang, L., Zheng, Y., Wang, C. and Zhao, Z. (2014) Morphological trend analysis of rice phytolith during the early Neolithic in the Lower Yangtze. Journal of Archaeological Science 49, 326–31.Google Scholar
Xie, S., Nott, C.J., Avsejs, L. A., Maddy, D., Chambers, F.M. and Evershed, R.P. (2004) Molecular and isotopic stratigraphy in an ombrotrophic mire for paleoclimate reconstruction. Geochimica et Cosmochimica Acta 68(13), 2849–62.Google Scholar
Yevdokimov, V.V. and Varfolomeev, V.V. (2002) Ehpoxa Bronzy Central’novo i Cevernovo Kazaxstana, Karaganda: E.A. Bukatov Karaganda State University.Google Scholar
Yohe, R.M. and Bamforth, D.B. (2013) Late Pleistocene protein residues from the Mahaffy cache, Colorado. Journal of Archaeological Science 40(5), 2337–43.Google Scholar
Young, D.L., Huyen, Y. and Allard, M.W. (1995) Testing the validity of the cytochrome b sequence from Cretaceous period bone fragments as dinosaur DNA. Cladistics 11(2), 199209.Google Scholar
Zaibert, V.F. (2009) Botaiskaya Kultura. Almaty: KazAkparat.Google Scholar
Zaibert, V.F., Tyulevaev, A., Zadorozhnyj, A.V. and Kulakov, U. (2007) Tajny Drevnyej Stepi: Issledovaniya Poseleniya Botaj (2004–2006). Kokshetau: Kokshetau University.Google Scholar
Zarky, A. (1976) Statistical analysis of site catchments at Ocos, Guatemala. In: Flannery, K.V. (ed.), The Early Mesoamerican Village. New York: Academic Press, pp. 117–28.Google Scholar
Zazzo, A., Bendrey, R., Vella, D., Moloney, A.P., Monahan, F.J. and Schmidt, O. (2012) A refined sampling strategy for intra-tooth stable isotope analysis of mammalian enamel. Geochimica et Cosmochimica Acta 84, 113.Google Scholar
Zazzo, A., Cerling, T.E., Ehleringer, J.R., Moloney, A.P., Monahan, F.J. and Schmidt, O. (2015) Isotopic composition of sheep wool records seasonality of climate and diet. Rapid Communications in Mass Spectrometry 29(15), 1357–69.Google Scholar
Zeder, M.A. (2012a) The broad spectrum revolution at 40: resource diversity, intensification, and an alternative to optimal foraging explanations. Journal of Anthropological Archaeology 31(3), 241–64.Google Scholar
Zeder, M.A. (2012b). The domestication of animals. Journal of Anthropological Research 68(2), 161–90.Google Scholar
Zeder, M.A. and Lapham, H.A. (2010) Assessing the reliability of criteria used to identify postcranial bones in sheep, Ovis, and goats, Capra. Journal of Archaeological Science 37(11), 2887–905.Google Scholar
Zhao, Z., Pearsall, D.M., Benfer, R.A. Jr. and Piperno, D. (1998) Distinguishing rice (Oryza sativa Poaceae) from wild Oryza species through phytolith analysis, II: finalised method. Economic Botany 52, 134–45.Google Scholar
Ziesemer, K.A., Mann, A.E., Sankaranarayanan, K., Schroeder, H., Ozga, A.T., Brandt, B.W., et al. (2015) Intrinsic challenges in ancient microbiome reconstruction using 16S rRNA gene amplification. Scientific Reports 5, 16498.Google Scholar
Zischler, H., Hoss, M., Handt, O., von Haeseler, A., van der Kuyl, A.C., Goudsmit, J. and Pääbo, S. (1995) Detecting dinosaur DNA. Science 268 (5214), 1192–3.Google Scholar
Zohary, D. and Hopf, M. (1988) Domestication of Plants in the Old World. Oxford: Clarendon.Google Scholar
Zvelebil, M. (1986) Mesolithic prelude and Neolithic revolution. In: Zvelebil, M. (ed.), Hunters in Transition: Mesolithic Societies of Temperate Eurasia and their Transition to Farming. Cambridge: Cambridge University Press, pp. 516.Google Scholar
Zvelebil, M. (1995) Hunting, gathering and husbandry? Management and food resources by the late Mesolithic communities of temperate Europe. In: Campana, D.V. (ed.), Before Farming: Hunter-Gatherer Society and Subsistence. Philadelphia, pa: MASCA, pp. 79104.Google Scholar
Zvelebil, M. (1998) Agricultural frontiers, Neolithic origins, and the transition to farming in the Baltic Basin. In: Zvelebil, M., Dennell, R. and Domańska, L. (eds), Harvesting the Sea, Farming the Forest: The Emergence of Neolithic Societies in the Baltic region. Sheffield: Sheffield Academic Press, pp. 928.Google Scholar
Zvelebil, M. and Rowley-Conwy, P. (1984) Transition to farming in northern Europe: a hunter-gatherer perspective. Norwegian Archaeological Review 17, 104–28.Google Scholar
Zvelebil, M. and Rowley-Conwy, P. (1986) Foragers and farmers in Atlantic Europe. In: Zvelebil, M. (ed.), Hunters in Transition: Mesolithic Societies of Temperate Eurasia and their Transition to Farming. Cambridge: Cambridge University Press, pp. 6793.Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

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

Available formats
×

Save book to Google Drive

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

Available formats
×