Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-4hhp2 Total loading time: 0 Render date: 2024-06-01T19:39:30.641Z Has data issue: false hasContentIssue false

Part II - Migratory Processes and Linguistic Dispersals between Yamnaya and the Corded Ware

Published online by Cambridge University Press:  29 April 2023

Kristian Kristiansen
Affiliation:
Göteborgs Universitet, Sweden
Guus Kroonen
Affiliation:
Universiteit Leiden
Eske Willerslev
Affiliation:
University of Copenhagen
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
The Indo-European Puzzle Revisited
Integrating Archaeology, Genetics, and Linguistics
, pp. 61 - 126
Publisher: Cambridge University Press
Print publication year: 2023

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

References

Allentoft, M. E., et al. 2015. Population genomics of Bronze Age Eurasia. Nature 522(7555): 167172.CrossRefGoogle ScholarPubMed
Andersen, S. T. 1993. History of vegetation and agriculture: At Hassing Huse Mose, Thy, Northwest Denmark, since the Ice Age. Journal of Danish Archaeology 11(1): 5779.CrossRefGoogle Scholar
Andersen, S. T. 1995. Pollen analytical investigations of barrows from the Funnel Beaker and Single Grave Cultures in the Vroue area, West Jutland, Denmark. Journal of Danish Archaeology 12(1): 107131.CrossRefGoogle Scholar
Andrades Valtueña, A., et al. 2017. The Stone Age plague and its persistence in Eurasia. Current Biology: CB 27(23): 36833691.e8.Google Scholar
Antanaitis-Jacobs, I., Daugnora, L., & Richards, M.. 2009. Diet in early Lithuanian prehistory and the new stable isotope evidence. Archaeologia Baltica 12: 1230.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: Princeton University Press.Google Scholar
Asam, T., Grupe, G., & Peters, J.. 2006. Menschliche Subsistenzstrategien im Neolithikum: Eine Isotopenanalyse bayerischer Skelettfunde. Anthropologischer Anzeiger; Bericht über die biologisch-anthropologische Literatur 64(1): 123.CrossRefGoogle Scholar
Bergemann, S. 2018. Zauschwitz (Landkreis Leipzig). Siedlungen und Gräber eines neolithischen Fundplatzes. Bonn: Dr. Rudolf Habelt GmbH.Google Scholar
Bourgeois, Q., & Kroon, E.. 2017. The impact of male burials on the construction of Corded Ware identity: Reconstructing networks of information in the 3rd millennium BC. PLoS ONE 12(10): e0185971.Google Scholar
Brandt, G., et al. 2013. Ancient DNA reveals key stages in the formation of Central European mitochondrial genetic diversity. Science 342(6155): 257261.CrossRefGoogle ScholarPubMed
Brandt, G., et al. 2015. Human paleogenetics of Europe: The known knowns and the known unknowns. Journal of Human Evolution 79: 7392.CrossRefGoogle ScholarPubMed
Brozio, J. P., et al. 2019. Monuments and economies: What drove their variability in the middle-Holocene Neolithic? The Holocene 29(10): 15581571.CrossRefGoogle Scholar
Burgess, C., & Shennan, S.. 1976. The Beaker phenomenon: Some suggestions. British Archaeological Reports.Google Scholar
Cassidy, L. M., et al. 2020. A dynastic elite in monumental Neolithic society. Nature 582(7812): 384388.CrossRefGoogle ScholarPubMed
Cramp, L. J. E. et al. 2014. Neolithic dairy farming at the extreme of agriculture in Northern Europe. Proceedings of the Royal Society B: Biological Sciences 281(1791): 20140819.CrossRefGoogle ScholarPubMed
Diers, S., et al. 2014. The Western Altmark versus Flintbek: Palaeoecological research on two megalithic regions. Journal of Archaeological Science 41: 185198.CrossRefGoogle Scholar
Diers, S. 2018. Mensch-Umweltbeziehungen zwischen 4000 und 2200 cal BC. Vegetationsgeschichtliche Untersuchungen an Mooren und trichterbecherzeitlichen Fundplätzen der Altmark. Bonn: Dr. Rudolf Habelt GmbH.Google Scholar
Diers, S., & Fritsch, B.. 2019. Changing environments in a megalithic landscape: The Altmark case. In: Müller, J., Hinz, M., & Wunderlich, M. (ed.), Megaliths – Societies – Landscapes. Early monumentality and social differentiation in Neolithic Europe, 719752. Bonn: Dr. Rudolf Habelt Verlag.Google Scholar
Doppler, T., et al. 2012. 14C-Datierung des endneolithischen Kollektivgrabes von Spreitenbach = Les datations radiocarbones de la sépulture collective de Spreitenbach. In: Doppler, Thomas & Alt, Kurt W. (ed.), Spreitenbach-Moosweg (Aargau, Schweiz) : ein Kollektivgrab um 2500 v.Chr. = Spreitenbach-Moosweg (Argovie, Suisse) : une sépulture collective vers 2500 av. J.-C., 85103. Basel: Archäologie Schweiz (Antiqua).Google Scholar
Doppler, T., et al. 2017. Landscape opening and herding strategies: Carbon isotope analyses of herbivore bone collagen from the Neolithic and Bronze Age lakeshore site of Zurich-Mozartstrasse, Switzerland. Quaternary International 436: 1828.CrossRefGoogle Scholar
Dörfler, W., et al. 2012. A high-quality annually laminated sequence from Lake Belau, Northern Germany: Revised chronology and its implications for palynological and tephrochronological studies. The Holocene 22(12): 14131426.Google Scholar
Feeser, I., et al. 2012. New insight into regional and local land-use and vegetation patterns in eastern Schleswig-Holstein during the Neolithic. In: M. Hinz & J. Müller (ed.), Siedlung, Grabenwerk, Grossteingrab. Studien zu Gesellschaft, Wirtschaft und Umwelt der Trichterbechergruppen im nördlichen Mitteleuropa. Frühe Monumentalität und Soziale Differenzierung, 159191. Bonn: Rudolf Habelt.Google Scholar
Feeser, I., et al. 2016. A mid-Holocene annually laminated sediment sequence from Lake Woserin: The role of climate and environmental change for cultural development during the Neolithic in Northern Germany. Holocene 26(6): 947963.Google Scholar
Feeser, I., et al. 2019. Human impact and population dynamics in the Neolithic and Bronze Age: Multi-proxy evidence from north-western Central Europe. The Holocene 29(10): 15961606.Google Scholar
Feeser, I., & Dörfler, W.. 2019. Land-use and environmental history at the Middle Neolithic settlement site Oldenburg-Dannau LA 77. Journal of Neolithic Archaeology 21: 157207.Google Scholar
Fernandes, D. M., et al. 2018. A genomic Neolithic time transect of hunter-farmer admixture in central Poland. Scientific Reports 8(1): 14879.Google Scholar
Filipović, D., et al. 2020. New AMS 14C dates track the arrival and spread of broomcorn millet cultivation and agricultural change in prehistoric Europe. Scientific Reports 10(1): 13698.CrossRefGoogle ScholarPubMed
Fischer, U. 1958. Mitteldeutschland und die Schnurkeramik. Ein kultursoziologischer Vergleich. Jahresschrift für mitteldeutsche Vorgeschichte 41/42: 254298.Google Scholar
Fornander, E. 2013. Dietary diversity and moderate mobility: Isotope evidence from Scanian Battle Axe Culture burials. Journal of Nordic Archaeological Science 18: 1329.Google Scholar
Fortunato, L., & Jordan, F.. 2010. Your place or mine? A phylogenetic comparative analysis of marital residence in Indo-European and Austronesian societies. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences 365(1559): 39133922.Google Scholar
Frieman, C. J., & Hofmann, D.. 2019. Present pasts in the archaeology of genetics, identity, and migration in Europe: A critical essay. World Archaeology 51(4): 528545.Google Scholar
Friis-Holm Egfjord, A., et al. 2021. Genomic steppe ancestry in skeletons from the Neolithic Single Grave culture in Denmark. PLoS ONE 16(1): e0244872.CrossRefGoogle ScholarPubMed
Furholt, M. 2014. Upending a “totality”: Re-evaluating Corded Ware variability in Late Neolithic Europe. Proceedings of the Prehistoric Society 80: 6786.Google Scholar
Furholt, M. 2018. Massive migrations? The impact of recent aDNA studies on our view of third millennium Europe. European Journal of Archaeology 21(2): 159191.Google Scholar
Furholt, M. 2019. Re-integrating archaeology: A contribution to aDNA studies and the migration discourse on the 3rd millennium BC in Europe. Proceedings of the Prehistoric Society 85: 115129.CrossRefGoogle Scholar
Furtwängler, A., et al. 2020. Ancient genomes reveal social and genetic structure of Late Neolithic Switzerland. Nature Communications 11(1): 1915.Google Scholar
Gamkrelidze, T. V., & Ivanov, V.. 1995. Indo-European and the Indo-Europeans: A reconstruction and historical analysis of a proto-language and proto-culture. Berlin: Mouton de Gruyter.CrossRefGoogle Scholar
Gerling, C., et al. 2017. High-resolution isotopic evidence of specialised cattle herding in the European Neolithic. PLoS ONE 12(7): e0180164.Google Scholar
Gimbutas, M. 1979. The three waves of Kurgan people into Old Europe, 4500–2500 BC. Archives Suisses d´anthropologie genérale 43(2): 113137.Google Scholar
Glob, P. V. 1944. Studier over den Jyske Enkeltgravskulturen. Aarbøger 1944: 1283.Google Scholar
Goldberg, A., et al. 2017. Ancient X chromosomes reveal contrasting sex bias in Neolithic and Bronze Age Eurasian migrations. Proceedings of the National Academy of Sciences of the United States of America 114(10): 26572662.Google Scholar
Goude, G., et al. 2019. A multidisciplinary approach to Neolithic life reconstruction. Journal of Archaeological Method and Theory 26(2): 537560.Google Scholar
Grossmann, R. 2016. Das dialektische Verhältnis von Schnurkeramik und Glockenbecher zwischen Rhein und Saale. Bonn: Verlag Dr. Rudolf Habelt GmbH.Google Scholar
Haak, W. 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 of the United States of America 105(47): 1822618231.CrossRefGoogle ScholarPubMed
Haak, W. et al. 2015. Massive migration from the steppe was a source for Indo-European languages in Europe. Nature 522(7555): 207211.Google Scholar
Harrison, R. J., & Heyd, V.. 2007. The transformation of Europe in the third millenium BC: The example of “Le Petit Chasseur I+III” (Sion, Valais, Switzerland). Praehistorische Zeitschrift 82: 129214.Google Scholar
Häusler, A. 1996. Invasionen aus der nordpontischen Steppen nach Mitteleuropa im Neolithikum und in der Bronzezeit: Realität oder Phantasieprodukt? Archäologische Informationen 19: 7588.Google Scholar
Hellman, S. 2008. Validating and testing the landscape reconstruction algorithm in southern Sweden: Towards quantitative reconstruction of past vegetation. Kalmar: School of Pure and Applied Natural Sciences, University of Kalmar.Google Scholar
Heron, C., et al. 2015. Cooking fish and drinking milk? Patterns in pottery use in the southeastern Baltic, 3300–2400 cal BC. Journal of Archaeological Science 63: 3343.CrossRefGoogle Scholar
Heyd, V. 2017. Kossinna’s smile. Antiquity 91(356): 348359.Google Scholar
Hinz, M., et al. 2012. Demography and the intensity of cultural activities: An evaluation of Funnel Beaker Societies (4200–2800 cal BC). Journal of Archaeological Science 39(10): 33313340.Google Scholar
Hübner, E. 2005. Jungneolithische Gräber auf der jütischen Halbinsel. Typologische und chronologische Studien zur Einzelgrabkultur (Nordiske Fortidsminder. Serie B). Copenhagen: Det Kongelige Oldskriftselskab.Google Scholar
Immel, A., et al. 2020. Gene-flow from steppe individuals into Cucuteni-Trypillia associated populations indicates long-standing contacts and gradual admixture. Scientific Reports 10(1): 4253.CrossRefGoogle ScholarPubMed
Ivanova, M. 2013. The Black Sea and the early civilizations of Europe, the Near East and Asia. Cambridge: Cambridge University Press.Google Scholar
Iversen, R. 2015. Sen tragtbæger- og enkeltgravskultur på de danske øer i sen mellemneolitikum, ca. 2850–2350 f. Kr. Strategi for yngre stenalders arkæologiske undersøgelser, November 2015: 58–82. Kulturstyrelsen.Google Scholar
Johannsen, N., & Laursen, S.. 2010. Routes and wheeled transport in late 4th–early 3rd millennium funerary customs of the Jutland peninsula: Regional evidence and European context. Praehistorische Zeitschrift 85(1): 1558.Google Scholar
Jones, E. R., et al. 2015. Upper Palaeolithic genomes reveal deep roots of modern Eurasians. Nature Communications 6: 8912.Google Scholar
Jones, E. R., et al. 2017. The Neolithic transition in the Baltic was not driven by admixture with early European farmers. Current Biology: CB 27(4): 576582.Google Scholar
Juras, A. et al. 2018. Mitochondrial genomes reveal an east to west cline of steppe ancestry in Corded Ware populations. Scientific Reports 8(1): 11603.Google Scholar
Kaiser, E. 2019. Das dritte Jahrtausend im osteuropäischen Steppenraum: Kulturhistorische Studien zu prähistorischer Subsistenzwirtschaft und Interaktion mit benachbarten Räumen (Berlin Studies of the Ancient World 37). Berlin: Edition Topoi.Google Scholar
Knipper, C., et al. 2017. Female exogamy and gene pool diversification at the transition from the Final Neolithic to the Early Bronze Age in central Europe. Proceedings of the National Academy of Sciences of the United States of America 114(38): 1008310088.Google Scholar
Knitter, D., et al. 2019. Transformations and site locations from a landscape archaeological perspective: The case of Neolithic Wagrien, Schleswig-Holstein, Germany. Land 8(4): 68.Google Scholar
Kossinna, G. 1910. Der Ursprung der Urfinnen und Urindogermanen und ihre Ausbreitung nach Osten. Mannus 1/2: 225245.Google Scholar
Kossinna, G. 1911. Die Herkunft der Germanen. Zur Methode der Siedlungsarchäologie. Bonn: Mannus-Bibliothek.Google Scholar
Krauß, R., et al. 2017. Chronology and development of the Chalcolithic necropolis of Varna I. Documenta Praehistorica 44: 282305.CrossRefGoogle Scholar
Kristiansen, K. 1989. Prehistoric migrations: The case of the Single Grave and Corded Ware Culture. Journal of Danish Archaeology 8: 211225.Google Scholar
Kristiansen, K., et al. 2017. Re-theorising mobility and the formation of culture and language among the Corded Ware Culture in Europe. Antiquity 91(356): 334347.Google Scholar
Lanting, J. N., Mook, W. G., & van der Waals, J. D.. 1973. C14 chronology and the Beaker problem. Helinium 13: 3858.Google Scholar
Lazaridis, I., & Reich, D.. 2017. Failure to replicate a genetic signal for sex bias in the steppe migration into central Europe. Proceedings of the National Academy of Sciences 114(20): E3873E3874.Google Scholar
Lechterbeck, J., et al. 2014. How was Bell Beaker economy related to Corded Ware and Early Bronze Age lifestyles? Archaeological, botanical and palynological evidence from the Hegau, Western Lake Constance region. Environmental Archaeology 19(2): 95113.Google Scholar
Linderholm, A. et al. 2020. Corded Ware cultural complexity uncovered using genomic and isotopic analysis from south-eastern Poland. Scientific Reports 10(1): 6885.CrossRefGoogle ScholarPubMed
Lipson, M. et al. 2017. Parallel palaeogenomic transects reveal complex genetic history of early European farmers. Nature 551(7680): 368372.Google Scholar
Mallory, J. P. 1989. In search of the Indo-Europeans: Language, archaeology and myth. London: Thames & Hudson.Google Scholar
Malmström, H., et al. 2019. The genomic ancestry of the Scandinavian Battle Axe Culture people and their relation to the broader Corded Ware horizon. Proceedings. Biological Sciences/The Royal Society 286(1912): 20191528.Google Scholar
Mathieson, I., et al. 2015. Genome-wide patterns of selection in 230 ancient Eurasians. Nature 528(7583): 499503.Google Scholar
Mathieson, I., et al. 2018. The genomic history of southeastern Europe. Nature 555: 197203.Google Scholar
Meller, H., et al. (ed). 2019. Late Neolithic and early Bronze Age settlement archaeology: 11th Archaeological Conference of Central Germany, October 18–20, 2018 in Halle (Saale). Halle: Landesamt für Denkmalpflege und Archäologie Sachsen-Anhalt, Landesmuseum für Vorgeschichte.Google Scholar
Mittnik, A., et al. 2018. The genetic prehistory of the Baltic Sea region. Nature Communications 9(1): 442.Google Scholar
Mittnik, A., et al. 2019. Kinship-based social inequality in Bronze Age Europe. Science 366(6466): 731734.Google Scholar
Monroy Kuhn, J. M., Jakobsson, M., & Günther, T.. 2018. Estimating genetic kin relationships in prehistoric populations. PLoS ONE 13(4): e0195491.Google Scholar
Müller, J. 2001. Zum Verhältnis von Schnurkeramik und jüngeren Trichterbechergruppen im Mittelelbe-Saale-Gebiet: Kontinuität oder Diskontinuität? In: Gohlisch, T. H. & Reisch, L. (ed.), Die Stellung der endneolithischen Chamer Kultur in ihrem räumlichen und zeitlichen Kontext (Kolloquien des Institutes für Ur- und Frühgeschichte Erlangen), 120136. Erlangen: Dr. Faustus.Google Scholar
Müller, J., et al. 2009. A revision of Corded Ware settlement pattern: New results from the Central European Low Mountain Range. Proceedings of the Prehistoric Society 75: 125142.Google Scholar
Müller, J. 2019. Boom and bust, hierarchy and balance: From landscape to social meaning: Megaliths and societies in Northern Central Europe. In: Müller, J., Hinz, M., & Wunderlich, M. (ed.), Megaliths – Societies – Landscapes. Early monumentality and social differentiation in Neolithic Europe, 2974. Bonn: Dr. Rudolf Habelt GmbH.Google Scholar
Müller, S. 1898. De jydske Enkeltgrave fra Stenalderen. Aarbøger 1898: 157282.Google Scholar
Münster, A., et al. 2018. 4000 years of human dietary evolution in central Germany, from the first farmers to the first elites. PLoS ONE 13(3): e0194862.Google Scholar
Narasimhan, V. M., et al. 2019. The formation of human populations in South and Central Asia. Science 365(6457).Google Scholar
Nielsen, P. O. 2020. The development of the two-aisled longhouse in the Neolithic and Early Bronze Age. In: Reedz-Sparrevohn, L., Thirup Kastholm, O., & Nielsen, P. O. (ed.), Houses for the living. Two-aisled houses from the Neolithic and Early Bronze Age in Denmark (The Royal Society of Northern Antiquities), 951. Copenhagen: University Press of Southern Denmark.Google Scholar
Odgaard, B. V. 1994. The Holocene vegetation history of northern West Jutland, Denmark. Nordic Journal of Botany 14(5): 546546.CrossRefGoogle Scholar
Olalde, I., et al. 2018. The Beaker phenomenon and the genomic transformation of northwest Europe. Nature 555: 190196.CrossRefGoogle ScholarPubMed
Piličiauskas, G., et al. 2017. The transition from foraging to farming (7000–500 cal BC) in the SE Baltic: A re-evaluation of chronological and palaeodietary evidence from human remains. Journal of Archaeological Science: Reports 14: 530542.Google Scholar
Piličiauskas, G., et al. 2018. The Corded Ware culture in the Eastern Baltic: New evidence on chronology, diet, beaker, bone and flint tool function. Journal of Archaeological Science: Reports 21: 538552.Google Scholar
Piličiauskas, G., et al. 2020. Fishers of the Corded Ware culture in the Eastern Baltic. Acta Archaeologica 91(1): 95120.Google Scholar
Racimo, F., et al. 2020. The spatiotemporal spread of human migrations during the European Holocene. Proceedings of the National Academy of Sciences of the United States of America 117(16): 89899000.CrossRefGoogle ScholarPubMed
Rascovan, N., et al. 2019. Emergence and spread of basal lineages of Yersinia pestis during the Neolithic decline. Cell 176(1/2): 295305.e10.Google Scholar
Rasmussen, P., & Olsen, J.. 2009. Soil erosion and land-use change during the last six millennia recorded in lake sediments of Gudme Sø, Fyn, Denmark. GEUS Bulletin 17: 3740.Google Scholar
Rasmussen, S., et al. 2015. Early divergent strains of Yersinia pestis in Eurasia 5,000 years ago. Cell 163(3): 571582.Google Scholar
Rasteiro, R., & Chikhi, L.. 2013. Female and male perspectives on the Neolithic transition in Europe: Clues from ancient and modern genetic data. PLoS ONE 8(4): e60944.Google Scholar
Renfrew, C. 1987. Archaeology and language: The puzzle of Indo-European origins. London: Pimlico.Google Scholar
Robb, J., & Harris, O. J. T.. 2018. Becoming gendered in European prehistory: Was Neolithic gender fundamentally different? American Antiquity 83(1): 128147.Google Scholar
Robson, H. K., et al. 2019. Diet, cuisine and consumption practices of the first farmers in the southeastern Baltic. Archaeological and Anthropological Sciences 11(8): 40114024.Google Scholar
Saag, L., et al. 2017. Extensive farming in Estonia started through a sex-biased migration from the steppe. Current Biology: CB 27(14): 21852193.e6.CrossRefGoogle ScholarPubMed
Saag, L., et al. 2021. Genetic ancestry changes in Stone to Bronze Age transition in the East European plain. Science Advances 7(4): eabd6535.Google Scholar
Salanova, L. 2011. Chronologie et facteurs d’évolution des sépultures individuelles campaniformes dans le Nord de la France. In: Salanova, Laure and Tcheremissinoff, Yaramila (ed.), Les sépultures individuelles campaniformes en France, 125142. Paris: CNRS Éditions.Google Scholar
Schrader, O. 1883. Sprachvergleichung und Urgeschichte: Linguistisch-historische Beiträge zur Erforschung des indogermanischen Altertums. Jena.Google Scholar
Schroeder, H., et al. 2019. Unraveling ancestry, kinship, and violence in a Late Neolithic mass grave. Proceedings of the National Academy of Sciences of the United States of America 116(22): 1070510710.CrossRefGoogle Scholar
Shennan, S., et al. 2013. Regional population collapse followed initial agriculture booms in mid-Holocene Europe. Nature Communications 4: 2486.Google Scholar
Siemen, P. 1997. Early Corded Ware culture. The A-horizon, fiction or fact? International symposium in Jutland, 2nd–7th May 1994 (Arkaeologiske Rapporter 2). Esbjerg: Esbjerg Museum.Google Scholar
Sjögren, K.-G. 2006. Ecology and economy in Stone Age and Bronze Age Scania (Skånska spår – Arkeologi längs Västkustbanan. National Heritage Board). Stockholm: Riksantikvarieämbetet.Google Scholar
Sjögren, K.-G., et al. 2020. Kinship and social organization in Copper Age Europe. A cross-disciplinary analysis of archaeology, DNA, isotopes, and anthropology from two Bell Beaker cemeteries. PLoS ONE 15(11): e0241278.Google Scholar
Sjögren, K.-G., Price, T. D., & Kristiansen, K.. 2016. Diet and mobility in the Corded Ware of Central Europe. PLoS ONE 11(5): e0155083.Google Scholar
Spyrou, M. A., et al. 2018. Analysis of 3800-year-old Yersinia pestis genomes suggests Bronze Age origin for bubonic plague. Nature Communications 9(1): 2234.Google Scholar
Spyrou, M. A., et al. 2019. Ancient pathogen genomics as an emerging tool for infectious disease research. Nature Reviews. Genetics 20(6): 323340.Google Scholar
Strahm, C. 2002. Tradition und Wandel der sozialen Strukturen vom 3. zum 2. vorchristlichen Jahrtausend. In: Müller, J. (ed.), Vom Endneolithikum zur Frühbronzezeit: Muster sozialen Wandels? (Tagung Bamberg 14.–16. Juni 2001) (Universitätsforschungen zur Prähistorischen Archäologie 90), 175194. Bonn: Habelt.Google Scholar
Strahm, C., & Buchvaldek, M.. 1991. Die kontinentaleuropäischen Gruppen der Kultur mit Schnurkeramik. Prague: Karolinum.Google Scholar
Struve, K. W. 1955. Die Einzelgrabkultur in Schleswig-Holstein. Neumünster: Offa-Bücher.Google Scholar
Sugita, S. 2007a. Theory of quantitative reconstruction of vegetation II: All you need is LOVE. The Holocene 17(2): 243257.Google Scholar
Sugita, S. 2007b. Theory of quantitative reconstruction of vegetation I: Pollen from large sites REVEALS regional vegetation composition. The Holocene 17(2): 229241.CrossRefGoogle Scholar
Szczepanek, A., et al. 2018. Understanding Final Neolithic communities in south-eastern Poland: New insights on diet and mobility from isotopic data. PLoS ONE 13(12): e0207748.Google Scholar
Szécsényi-Nagy, A., et al. 2015. Tracing the genetic origin of Europe’s first farmers reveals insights into their social organization. Proceedings. Biological Sciences/The Royal Society 282(1805).CrossRefGoogle ScholarPubMed
Theuerkauf, M., et al. 2016. A matter of dispersal: REVEALSinR introduces state-of-the-art dispersal models to quantitative vegetation reconstruction. Vegetation History and Archaeobotany 25(6): 541553.CrossRefGoogle Scholar
Theuerkauf, M., & Couwenberg, J.. 2018. ROPES reveals past land cover and PPEs from single pollen records. Frontiers of Earth Science in China 6(14).Google Scholar
Ullrich, M. 2008. Endneolithische Siedlungskeramik aus Ergersheim, Mittelfranken. Untersuchungen zur Chronologie von Schnurkeramik- und Glockenbechern an Rhein, Main und Neckar (Universitätsforschungen zur Prähistorischen Archäologie). Bonn: Habelt.Google Scholar
Veit, U. 1989. Ethnic concepts in German prehistory: A case study on the relationship between cultural identity and archaeological objectivity. In: Shennan, S. (ed.), Archaeological approaches to cultural identity. London: Routledge.Google Scholar
Wang, C.-C., et al. 2019. Ancient human genome-wide data from a 3000-year interval in the Caucasus corresponds with eco-geographic regions. Nature Communications 10(1): 590.CrossRefGoogle ScholarPubMed
Wentink, K. 2020. Stereotype. The role of grave sets in Corded Ware and Bell Beaker funerary practices. Leiden: Sidestone Press.Google Scholar
Werens, K., Szczepanek, A., & Jarosz, P.. 2018. Light stable isotope analysis of diet in Corded Ware culture communities: Święte, Jarosław District, South-Eastern Poland. Baltic-Pontic Studies, 23(1): 229245.Google Scholar
Wiethold, J. 1998. Studien zur jüngeren postglazialen Vegetations- und Siedlungsgeschichte im östlichen Schleswig-Holstein (Universitätsforschungen zur Prähistorischen Archäologie). Bonn: Habelt.Google Scholar
Wilkins, J. F., & Marlowe, F. W.. 2006. Sex-biased migration in humans: What should we expect from genetic data? BioEssays 28(3): 290300.Google Scholar
Yu, H., et al. 2020. Paleolithic to Bronze Age Siberians reveal connections with First Americans and across Eurasia. Cell 181(6): 12321245.e20.Google Scholar
Zanon, M., et al. 2018. European forest cover during the past 12,000 years: A palynological reconstruction based on modern analogs and remote sensing. Frontiers in Plant Science 9: 253253.Google Scholar

References

Allentoft, M. E., Sikora, M., Sjögren, K.-G., … & Willerslev, E.. 2015. Population genomics of Bronze Age Eurasia. Nature 522(7555): 167172.Google Scholar
Anthony, D. W., & Brown, D. R.. 2017. The dogs of war: A Bronze Age initiation ritual in the Russian steppes. Journal of Anthropological Archaeology 48: 134148.Google Scholar
Barraud, C., & Platenkamp, J. D. M.. 1990. Rituals and the comparison of societies. Bijdragen Tot Taal, Land En Volkenkunde KITLV 146: 103123.Google Scholar
Bourgeois, Q. P. J. 2013. Monuments on the horizon: The formation of the barrow landscape throughout the 3rd and 2nd millennium BC. Leiden: Sidestone Press.Google Scholar
Bourgeois, Q. P. J., & Kroon, E. J.. 2017. The impact of male burials on the construction of Corded Ware identity: Reconstructing networks of information in the 3rd millennium BC. PLoS ONE 12(10): e0185971.Google Scholar
Brandt, G., Haak, W., Adler, C. J., & Alt, K. W.. 2013. Ancient DNA reveals key stages in the formation of Central European mitochondrial genetic diversity. Science 342(6155): 257261.Google Scholar
Buchvaldek, M. 1986. Zum gemeineuropäischen Horizont der Schnurkeramik. Praehistorische Zeitschrift 61(2): 129151.CrossRefGoogle Scholar
Centola, D., & Baronchelli, A.. 2015. The spontaneous emergence of conventions: An experimental study of cultural evolution. Proceedings of the National Academy of Sciences 112(7): 201418838.Google Scholar
Centola, D., Gonzalez-Avella, J. C., Eguiluz, V. M., & San Miguel, M.. 2007. Homophily, cultural drift, and the co-evolution of cultural groups. Journal of Conflict Resolution 51(6): 905929.Google Scholar
Drenth, E., & Lohof, E.. 2005. Mounds for the dead: Funerary and burial ritual in Beaker period, Early and Middle Bronze Age. In Louwe Kooijmans, L. P., van den Broeke, P. W., Fokkens, H., & van Gijn, A. L. (ed.), The prehistory of the Netherlands, 433454. Amsterdam: Amsterdam University Press.Google Scholar
Furholt, M. 2014. Upending a “totality”: Re-evaluating Corded Ware variability in Late Neolithic Europe. Proceedings of the Prehistoric Society 80: 120.Google Scholar
Furholt, M. 2017. Massive migrations? The impact of recent aDNA studies on our view of third millennium Europe. European Journal of Archaeology 21(2): 159191.CrossRefGoogle Scholar
Goldberg, A., Günther, T., Rosenberg, N.A., & Jakobsson, M.. 2017. Ancient X chromosomes reveal contrasting sex bias in Neolithic and Bronze Age Eurasian migrations. Proceedings of the National Academy of Sciences of the United States of America 114(10): 26572662.Google Scholar
Haak, W., Lazaridis, I., Patterson, N., … & Reich, D.. 2015. Massive migration from the steppe was a source for Indo-European languages in Europe. Nature 522(7555): 207211.Google Scholar
Heyd, V. 2017. Kossinna’s smile. Antiquity 91(356): 348359.Google Scholar
Hübner, E. 2005. Jungneolithischen Gräber auf der Jütischen Halbinsel; Typologische und chronologische Studien zur Einzelgrabkultur. Copenhagen: Det Kongeliche Nordiske Oldskriftselskab.Google Scholar
Juras, A., Chyleński, M., Ehler, E., … & Kośko, A.. 2018. Mitochondrial genomes reveal an east to west cline of steppe ancestry in Corded Ware populations. Scientific Reports 8(1): 11603.CrossRefGoogle ScholarPubMed
Kershaw, K. 2000. The one-eyed god: Odin and the (Indo-)Germanic Männerbund (Journal of Indo-European Studies Monographs 36). Washington, DC: Institute for the Study of Man.Google Scholar
Knipper, C., Mittnik, A., Massy, K., … & Stockhammer, P. W.. 2017. Female exogamy and gene pool diversification at the transition from the Final Neolithic to the Early Bronze Age in central Europe. Proceedings of the National Academy of Sciences 114(38): 1008310088.Google Scholar
Kristiansen, K., Allentoft, M. E., Frei, K. M., … & Willerslev, E.. 2017. Re-theorising mobility and the formation of culture and language among the Corded Ware Culture in Europe. Antiquity 91(356): 334347.Google Scholar
Lave, J., & Wenger, E.. 1991. Situated learning: Legitimate peripheral participation. Cambridge: Cambridge University Press.Google Scholar
Metcalf, P., & Huntington, R.. 1991. Celebrations of death: The anthropology of mortuary ritual. Cambridge: Cambridge University Press.Google Scholar
Oestigaard, T., & Goldhahn, J.. 2006. From the dead to the living: Death as transactions and re-negotiations. Norwegian Archaeological Review 39(1): 2748.CrossRefGoogle Scholar
Olalde, I., Brace, S., Allentoft, M. E., … & Reich, D.. 2018. The Beaker phenomenon and the genomic transformation of northwest Europe. Nature 555(7695): 190196.CrossRefGoogle ScholarPubMed
Pospieszny, Ł., Sobkowiak-Tabaka, I., Price, T. D., … & Winiarska-Kabacińska, M.. 2015. Remains of a late Neolithic barrow at Kruszyn. A glimpse of ritual and everyday life in early Corded Ware societies of the Polish Lowland. Praehistorisch Zeitschrift 90(1–2): 185213.Google Scholar
Price, T. D., Knipper, C., Grupe, G., & Smrcka, V.. 2004. Strontium isotopes and prehistoric human migration: The Bell Beaker period in Central Europe. European Journal of Archaeology 7(1): 940.Google Scholar
Salton, G., & McGill, M. J.. 1983. Introduction to modern information retrieval. New York: McGraw-Hill Book Company.Google Scholar
Sjögren, K.-G., Price, T. D., & Kristiansen, K.. 2016. Diet and mobility in the Corded Ware of Central Europe. PLoS ONE 11(5): e0155083.Google Scholar
Smejda, L., Turek, J., & Thrane, H.. 2006. Archaeology of burial mounds. Plzen: University of West-Bohemia, Department of Archaeology.Google Scholar
Sørensen, M. L. S. 2015. “Paradigm lost”: On the state of typology within archaeological theory. In Kristiansen, K., Šmejda, L., & Turek, J. (ed.), Paradigm found, archaeological theory present, past and future, 8494. Oxford: Oxbow Books.Google Scholar
Turek, J. 2017. Sex, transsexuality and archaeological perception of gender identities. Archaeologies: Journal of the World Archaeological Congress 12(3): 340358.Google Scholar
Van Giffen, A. E. 1935. Twee grafheuvels te Nieuw Roden, Gem. Roden. Oudheidkundige Aantekeningen over Drenthse Vondsten 2: 117–8.Google Scholar
Vander Linden, M. 2004. Polythetic networks, coherent people: A new historical hypothesis for the Bell Beaker phenomenon. In Czebreszuk, J. (ed.), Similar but different; Bell Beakers in Europe, 3562. Leiden: Sidestone Press.Google Scholar
vander Linden, M. 2016. Population history in third-millennium-BC Europe: Assessing the contribution of genetics. World Archaeology 48(5): 714728.Google Scholar
Wenger, E. 1998. Communities of practice: Learning, meaning, and identity. Cambridge: Cambridge University Press.Google Scholar

References

Adams, Douglas Q. 2013. A dictionary of Tocharian B. 2nd ed., revised and greatly enlarged (Leiden Studies in Indo-European 10). Amsterdam & New York: Rodopi.Google Scholar
Anthony, David W. 2007. The horse, the wheel, and language: How Bronze-Age riders from the Eurasian steppes shaped the modern world. Princeton (NJ) & Oxford: Princeton University Press.Google Scholar
Atkinson, Quentin D., & Gray, Russell D.. 2006. Are accurate dates an intractable problem for historical linguistics. In: Lipo, C. P. (ed.), Mapping our ancestors: Phylogenetic approaches in anthropology and prehistory, 269296. New Brunswick (NJ): Aldine Transaction.Google Scholar
Beekes, Robert S. P. 2010. Etymological dictionary of Greek (Leiden Indo-European Etymological Dictionary Series 10). Leiden & Boston: Brill.Google Scholar
Bjørn, Rasmus Gudmundsen. 2017. Foreign elements in the Proto-Indo-European vocabulary: A comparative loanword study. MA thesis, University of Copenhagen.Google Scholar
Blažek, Václav. 2017. Indo-European “gold” in time and space. Journal of Indo-European Studies 45(3/4): 267311.Google Scholar
Bouckaert, Remco, Lemey, Philippe, Dunn, Michael, Greenhill, Simon J., Alekseyenko, Alexander V., Drummond, Alexei J., Gray, Russell D., Suchard, Marc A., & Atkinson, Quentin D.. 2012. Mapping the origins and expansion of the Indo-European language family. Science 337: 957960.Google Scholar
Bouckaert, Remco, Lemey, Philippe, Dunn, Michael, Greenhill, Simon J., Alekseyenko, Alexander V., Drummond, Alexei J., Gray, Russell D., Suchard, Marc A., & Atkinson, Quentin D.. 2013. Corrections and clarifications. Science 342: 1446.Google Scholar
Boutkan, Dirk, & Kossmann, M.. 2001. On the etymology of “silver.” North-Western European Language Evolution 50: 511.Google Scholar
Campbell, Lyle, & Poser, William J.. 2008. Language classification: History and method. Cambridge: Cambridge University Press.Google Scholar
Chakrabarti, Dilip K. 1979. Iron in early Indian literature. The Journal of the Royal Asiatic Society of Great Britain and Ireland 1979(1): 2230.Google Scholar
Chang, Will, Cathcart, Chundra, Hall, David, & Garrett, Andrew. 2015. Ancestry-constrained phylogenetic analysis supports the Indo-European steppe hypothesis. Language 91(1): 194244.Google Scholar
Coleman, Robert. 1988. Review of Colin Renfrew, Archaeology and language: The puzzle of Indo-European origins (New York: Cambridge University Press, 1987). Current Anthropology 29(3): 449–453.Google Scholar
de Vaan, Michiel. 2008. Etymological dictionary of Latin and the other Italic languages (Leiden Indo-European Etymological Dictionary Series 7). Leiden & Boston: Brill.Google Scholar
Dressler, Wolfgang U. 1965. Methodische Vorfragen bei der Bestimmung der “Urheimat.” Die Sprache 11(1/2): 2560, 217.Google Scholar
Driessen, C. Michiel. 2003. *h2é‑h2us‑o‑, the Proto-Indo-European term for ‘gold’. Journal of Indo-European Studies 31: 347362.Google Scholar
Ehret, Christopher. 2015. Agricultural origins: What linguistic evidence reveals. In: Barker, Graeme & Goucher, C. (ed.), The Cambridge world history. Vol. 2. A world with agriculture, 5592. Cambridge: Cambridge University Press.Google Scholar
Frisk, Hjalmar. 1970. Griechisches etymologisches Wörterbuch. Vol. 2. Κρ–Ω. Heidelberg: Winter.Google Scholar
Gray, Russell D., & Atkinson, Quentin D.. 2003. Language-tree divergence times support the Anatolian theory of Indo-European origin. Nature 426: 435439.Google Scholar
Greenhill, Simon J., Heggarty, Paul, & Gray, Russell D.. 2021. Bayesian phylolinguistics. In: Janda, Richard D., Joseph, Brian D., & Vance, Barbara S. (ed.), The handbook of historical linguistics. Vol. 2, 226253. Hoboken (NJ): John Wiley & Sons.Google Scholar
Heggarty, Paul. 2015. Prehistory through language and archaeology. In: Bowern, Claire & Evans, Bethwyn (ed.), The Routledge handbook of historical linguistics, 598626. Oxford & New York: Routledge.Google Scholar
Heggarty, Paul. 2021. Cognacy databases and phylogenetic research on Indo-European. Annual Review of Linguistics 7: 371394.Google Scholar
Hock, Hans Henrich, & Joseph, Brian D.. 2009. Language history, language change, and language relationship: An introduction to historical and comparative linguistics. 2nd. revised edition (Trends in Linguistics. Studies and Monographs 218). Berlin & New York: de Gruyter.Google Scholar
Holm, Hans J. 2008. The distribution of data in word lists and its impact on the subgrouping of languages. In: Preisach, Christine, Burkhardt, Hans, Schmidt-Thieme, Lars, & Decker, Reinhold (ed.), Data analysis, machine learning and applications: Proceedings of the 31st Annual Conference of the Gesellschaft für Klassifikation e.V., Albert-Ludwigs-Universität Freiburg, March 7–9, 2007, 629636. Berlin & Heidelberg: Springer.CrossRefGoogle Scholar
Huld, Martin E. 2012. Some observations on the development of Indo-European metallurgy. In: Huld, Martin E., Jones-Bley, Karlene, & Miller, Dean (ed.), Archaeology and language: Indo-European studies presented to James P. Mallory (Journal of Indo-European Studies, Monograph Series 60), 281356. Washington, DC: Institute for the Study of Man.Google Scholar
Huld, Martin E., & Mallory, James P.. 1997a. Iron. In: Mallory, James P. & Adams, Douglas Q. (ed.), Encyclopedia of Indo-European culture, 313314. London & Chicago: Fitzroy Dearborn.Google Scholar
Huld, Martin E., & Mallory, James P.. 1997b. Metal. In: Mallory, James P. & Adams, Douglas Q. (ed.), Encyclopedia of Indo-European culture, 379380. London & Chicago: Fitzroy Dearborn.Google Scholar
Huld, Martin E., & Mallory, James P.. 1997c. Silver. In: Mallory, James P. & Adams, Douglas Q. (ed.), Encyclopedia of Indo-European culture, 518519. London & Chicago: Fitzroy Dearborn.Google Scholar
Joki, Aulis J. 1973. Uralier und Indogermanen: die älteren Berührungen zwischen den uralischen und indogermanischen Sprachen. Helsinki: Suomalais-Ugrilainen Seura.Google Scholar
Kallio, Petri. 2004. Tocharian loanwords in Samoyed? In: Hyvärinen, Irma, Kallio, Petri, & Korhonen, Jarmo (ed.), Etymologie, Entlehnungen und Entwicklungen: Festschrift für Jorma Koivulehto zum 70. Geburtstag (Mémoires de la Société Néophilologique 63), 129137. Helsinki: Société Néophilologique.Google Scholar
Kloekhorst, Alwin. 2008. Etymological dictionary of the Hittite inherited lexicon (Leiden Indo-European Etymological Dictionary Series 5). Leiden & Boston: Brill.Google Scholar
Kroonen, Guus. 2013. Etymological dictionary of Proto-Germanic (Leiden Indo-European Etymological Dictionary Series 11). Leiden & Boston: Brill.Google Scholar
Kümmel, Martin Joachim. 2017. Even more traces of the accent in Armenian? The development of tenues after sonorants. In: Hansen, Bjarne Simmelkjær Sandgaard, Hyllested, Adam, Jørgensen, Anders Richardt, Kroonen, Guus, Larsson, Jenny Helena, Whitehead, Benedicte Nielsen, Olander, Thomas, & Søborg, Tobias Mosbæk (ed.), Usque ad radices: Indo-European studies in honour of Birgit Anette Olsen (Copenhagen Studies in Indo-European), 439452. Copenhagen: Museum Tusculanum.Google Scholar
Lehmann, Winfred P. 1986. A Gothic etymological dictionary. Leiden: Brill.Google Scholar
Mallory, James P. 2018. Linguistic palaeontology. Paper presented at Languages and Migrations in Prehistoric Europe: Roots of Europe Summer Seminar, National Museum of Denmark and the University of Copenhagen, August 7–12, 2018.Google Scholar
Mallory, James P. 2019. Proto-Indo-European, Proto-Uralic, and Nostratic: A brief excursus into the comparative study of proto-languages. In: Olsen, Birgit Anette, Olander, Thomas, & Kristiansen, Kristian (ed.), Tracing the Indo-Europeans: New evidence from archaeology and historical linguistics, 3558. Oxford & Philadelphia: Oxbow.Google Scholar
Mallory, James P., & Adams, Douglas Q.. 2006. The Oxford introduction to Proto-Indo-European and the Proto-Indo-European world. Oxford & New York: Oxford University Press.Google Scholar
Mallory, James P., & Huld, Martin E.. 1984. Proto-Indo-European ‘silver’. Zeitschrift für vergleichende Sprachforschung 97: 112.Google Scholar
Martirosyan, Hrach K. 2010. Etymological dictionary of the Armenian inherited lexicon (Leiden Indo-European Etymological Dictionary Series 8). Leiden & Boston: Brill.CrossRefGoogle Scholar
Matasović, Ranko. 2009. Etymological dictionary of Proto-Celtic (Leiden Indo-European Etymological Dictionary Series 9). Leiden & Boston: Brill.Google Scholar
Mayrhofer, Manfred. 1986. Etymologisches Wörterbuch des Altindoarischen. Vol. 1. Heidelberg: Winter.Google Scholar
Mayrhofer, Manfred. 1996. Etymologisches Wörterbuch des Altindoarischen. Vol. 2. Heidelberg: Winter.Google Scholar
Meillet, Antoine. 1903. Introduction à l’étude comparative des langues indo-européennes. Paris: Hachette.Google Scholar
Melchert, H. Craig. 1994. Anatolian historical phonology (Leiden Studies in Indo-European 3). Amsterdam & Atlanta: Rodopi.Google Scholar
Mühlenbach, Karl, & Endzelīns, Jānis. 1923–1925. Lettisch-deutsches Wörterbuch. Redigiert, ergänzt und fortgesetzt von J. Endzelin. Vol. 1. Riga: Izglītības ministrija; Kultūras fonds.Google Scholar
Nakhleh, Luay, Ringe, Donald A., & Warnow, Tandy. 2005. Perfect phylogenetic networks: A new methodology for reconstructing the evolutionary history of natural languages. Language 81: 382420.Google Scholar
Neumann, Günther. 1975. Frühe Indogermanen und benachbarte Sprachgruppen. In: Karl J. Narr (ed.), Handbuch der Urgeschichte. Vol. 2. Jüngere Steinzeit und Steinkupferzeit. Frühe Bodenbau- und Viehzuchtkulturen, 673–689. Bern & Munich: Francke.Google Scholar
Nichols, Johanna. 1990. Linguistic diversity and the first settlement of the New World. Language 66(3): 475521.Google Scholar
Nichols, Johanna. 1997. The epicentre of the Indo-European linguistic spread. In: Blench, Roger & Spriggs, Matthew (ed.), Archaeology and language. Vol. 1. Theoretical and methodological orientations (One World Archaeology 27), 122148. London & New York: Routledge.Google Scholar
Olander, Thomas. 2017. Drinking beer, smoking tobacco and reconstructing prehistory. In: Hansen, Bjarne Simmelkjær Sandgaard, Hyllested, Adam, Jørgensen, Anders Richardt, Kroonen, Guus, Larsson, Jenny Helena, Whitehead, Benedicte Nielsen, Olander, Thomas, & Søborg, Tobias Mosbæk (ed.), Usque ad radices: Indo-European studies in honour of Birgit Anette Olsen (Copenhagen Studies in Indo-European 8), 605618. Copenhagen: Museum Tusculanum.Google Scholar
Olander, Thomas. 2018. Connecting the dots: The Indo-European family tree as a heuristic device. In: Goldstein, David, Jamison, Stephanie, & Vine, Brent (ed.), Proceedings of the 29th UCLA Indo-European Conference, 181202. Bremen: Hempen.Google Scholar
Olander, Thomas. 2019a. The Indo-European homeland: Introducing the problem. In: Olsen, Birgit Anette, Olander, Thomas, & Kristiansen, Kristian (ed.), Tracing the Indo-Europeans: New evidence from archaeology and historical linguistics, 734. Oxford & Philadelphia: Oxbow.Google Scholar
Olander, Thomas. 2019b. Indo-European cladistic nomenclature. Indogermanische Forschungen 124: 231244.CrossRefGoogle Scholar
Olsen, Birgit Anette. 1999. The noun in Biblical Armenian: Origin and word-formation – with special emphasis on the Indo-European heritage (Trends in Linguistics. Studies and Monographs 119). Berlin & New York: Mouton de Gruyter.Google Scholar
Orël, Vladimir E. 1998. Albanian etymological dictionary. Leiden, Boston, & Cologne: Brill.Google Scholar
Pereltsvaig, Asya, & Lewis, Martin W.. 2015. The Indo-European controversy: Facts and fallacies in historical linguistics. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Peyrot, Michaël. 2019. Indo-Uralic, Indo-Anatolian, Indo-Tocharian. In: Kloekhorst, Alwin & Pronk, Tijmen (ed.), The precursors of Proto-Indo-European: The Indo-Anatolian and Indo-Uralic hypotheses (Leiden Studies in Indo-European 21), 186202. Leiden & Boston: Brill.Google Scholar
Prósper, Blanca María. 2011. The Hispano-Celtic divinity ILVRBEDA, gold mining in Western Hispania and the syntactic context of Celtiberian arkatobezom ‘silver mine’. Die Sprache 49(1): 5383.Google Scholar
Puhvel, Jaan. 1991. Hittite etymological dictionary. Vol. 3. Words beginning with H. Berlin & New York: Mouton de Gruyter.Google Scholar
Rasmussen, Jens Elmegård. 1989. Studien zur Morphophonemik der indogermanischen Grundsprache (Innsbrucker Beiträge zur Sprachwissenschaft 55). Innsbruck: Institut für Sprach-wissenschaft der Universität Innsbruck.Google Scholar
Rau, Wilhelm. 1974. Metalle und Metallgeräte im vedischen Indien (Abhandlungen der Akademie der Wissenschaften und der Literatur. Geistes- und sozialwissenschaftliche Klasse, Jahrgang 1973, 8). Mainz: Akademie der Wissenschaften und der Literatur.Google Scholar
Ringe, Donald A. 1998. Some consequences of a new proposal for subgrouping the IE family. In: Bergen, Benjamin K., Madelaine, C. Plauché, , & Bailey, Ashlee C. (ed.), Proceedings of the Twenty-Fourth Annual Meeting of the Berkeley Linguistics Society, February 14–16, 1998. Special session on Indo-European subgrouping and internal relations, 3246. Berkeley: Berkeley Linguistics Society.Google Scholar
Ringe, Donald A. 2022. What we can (and can’t) learn from computational cladistics. In: Olander, Thomas (ed.), The Indo-European languages: A phylogenetic perspective, 5262. Cambridge: Cambridge University Press.Google Scholar
Ringe, Donald A., & Eska, Joseph F.. 2013. Historical linguistics: Toward a twenty-first century reintegration. Cambridge etc.: Cambridge University Press.Google Scholar
Ringe, Donald A., Warnow, Tandy, & Taylor, Ann. 2002. Indo-European and computational linguistics. Transactions of the Philological Society 100(1): 59129.Google Scholar
Rix, Helmut, Kümmel, Martin, Zehnder, Thomas, Lipp, Reiner, & Schirmer, Brigitte (ed.). 2001. Lexikon der indogermanischen Verben. 2nd expanded and improved edition. Wiesbaden: Reichert.Google Scholar
Schrader, Otto. 1890. Sprachvergleichung und Urgeschichte: Linguistisch-historische Beiträge zur Erforschung des indogermanischen Altertums. Dritte neubearbeitete Auflage. Jena: Costenoble.Google Scholar
Schrijver, Peter. 1991. The reflexes of the Proto-Indo-European laryngeals in Latin (Leiden Studies in Indo-European 2). Amsterdam & Atlanta: Rodopi.Google Scholar
Skjærvø, Prods Oktor. 1995. The Avesta as source for the early history of the Iranians. In: Erdosy, George (ed.), The Indo-Aryans of ancient South Asia: Language, material culture and ethnicity (Indian Philology and South Asian Studies 1), 155176. Berlin & New York: De Gruyter.Google Scholar
Tylecote, R. F. 1992. A history of metallurgy. 2nd ed. London: Maney.Google Scholar
Vanséveren, Sylvie. 2012. Noms de métaux dans les textes hittites. Anatolica 38: 203219.Google Scholar
Vasmer, Max. 1986. Ėtimologičeskij slovar’ russkogo jazyka [Etymological dictionary of the Russian language]. Vol. 2. E–Muž. 2nd ed. Translated with additions by O. N. Trubačëv. Moskow: Progress.Google Scholar
Vennemann, Theo. 1997. Some West Indo-European words of uncertain origin. In: Hickey, Raymond & Puppe, Stanislav (ed.), Language history and linguistic modelling. A Festschrift for Jacek Fisiak (Trends in Linguistics. Studies and Monographs 101), 879908. Berlin & New York: Mouton de Gruyter.Google Scholar
Walde, Alois, & Hofmann, J. B.. 1938. Lateinisches etymologisches Wörterbuch. 3., neubearbeitete Auflage von J. B. Hofmann (Indogermanische Bibliothek, Abt. 1, Reihe 2, Band 1). Heidelberg: Winter.Google Scholar
Weiss, Michael. 2020. Outline of the historical and comparative grammar of Latin. 2nd ed. Ann Arbor (MI) & New York: Beech Stave.Google Scholar
Witczak, Krzysztof Tomasz. 1990. Tocharian A nkiñc, B ñkante ‘silver’. Tocharian and Indo-European Studies 4: 4748.Google Scholar
Witzel, Michael. 2005. Indocentrism: Autochthonous visions of ancient India. In: Bryant, Edwin & Patton, Laurie L. (ed.), The Indo-Aryan controversy: Evidence and inference in Indian history, 341404. London & New York: Routledge.Google Scholar
Yamazaki, Yoko. 2009. The Saussure effect in Lithuanian. The Journal of Indo-European Studies 37(3): 430461.Google Scholar

References

Abaev = V.I. Abaev. 1958–89. Istoriko-etimologičeskij slovar’ osetinskogo jazyka. 4 vols. Leningrad: Nauka.Google Scholar
Adams, Douglas Q. 2013. A dictionary of Tocharian B. Amsterdam & New York: Rodopi.Google Scholar
Aikio, Ante. 2015. The Finnic ‘secondary e-stems’ and Proto-Uralic vocalism. Journal de la Société Finno-Ougrienne 95: 2566.Google Scholar
Akanuma, Hideo. 2007. Analysis of iron and copper production activity in central Anatolia during the Assyrian colony period. Anatolian Archaeological Studies 16: 125139.Google Scholar
Alessio, Giovanni. 1941. L’etrusco e due problemi etimologici latini. Aevum 15(4): 545558.Google Scholar
d’Arbois de Jubainville, Henri. 1902. Cours de litterature celtique XII: Principaux auteurs de l’antiquité à consulter sur l’histoire des Celtes depuis les temps les plus anciens jusqu’au règne de Theodose Ier. Paris: Fontemoing.Google Scholar
Bailey, Harold W. 1979. Dictionary of Khotan Saka. Cambridge: Cambridge University Press.Google Scholar
Baxter-Sagart = Baxter, William H. & Sagart, Laurent. 2014–. List of Old Chinese reconstructions by Grammata Seria Recensa number. http://ocbaxtersagart.lsait.lsa.umich.edu/.Google Scholar
Bebermeier, Wiebke et al. 2016. The coming of iron in a comparative perspective. In: Graßhoff, Gerd & Meyer, Michael (ed.), Space and knowledge: Topoi Research Group articles, 152189. Berlin: Excellence Cluster Topoi.Google Scholar
Beckwith, Miles. 1998. The ‘Hanging of Hera’ and the meaning of Greek ἄκμων. Harvard Studies in Classical Philology 98: 91102.Google Scholar
Beekes, Robert S. P. 1999. The Greek word for ‘lead’. Münchener Beiträge zur Sprachwissenschaft 59: 714.Google Scholar
Beekes, Robert S. P. 2002. The prehistory of the Lydians, the origin of the Etruscans, Troy and Aeneas. Bibliotheca Orientalis 59: 205241.Google Scholar
Bejko, Lorenc, Fenton, Todd, & Foran, David. 2006. Recent advances in Albanian mortuary archaeology, human osteology, and ancient DNA. In: Bejko, Lorenc & Hodges, Richard (ed.), New directions in Albanian archaeology: Studies presented to Muzafer Korkuti, 309322. Tirana: International Centre for Albanian Archaeology.Google Scholar
Belardi, Walter. 1949. Review of Traité de phonétique grecque by Michel Lejeune (Paris: Klincksieck, 1947). Maia 2: 308312.Google Scholar
Bellamy, Kate. 2018. Investigating interaction between South America and West Mexico through the lexicon of metallurgy. In: Iversen, Rune & Kroonen, Guus (ed.), Digging for words: Archaeolinguistic case studies from the XV Nordic TAG Conference held at the University of Copenhagen, 16–18 April 2015, 119. Oxford: BAR Publishing.Google Scholar
Benzing, Johannes. 1983. Chwaresmischer Wortindex. Wiesbaden: Harrassowitz.Google Scholar
Berger, D. et al. 2019. Isotope systematics and chemical composition of tin ingots from Mochlos (Crete) and other Late Bronze Age sites in the eastern Mediterranean Sea: An ultimate key to tin provenance? PLoS ONE 14(6): e0218326.Google Scholar
Blažek, Václav. 2010. Indo-European “smith” and his divine colleagues (JIES Monograph No. 58). Washington, DC: Institute for the Study of Man.Google Scholar
Blažek, Václav. 2017. Indo-European “gold” in time and space. Journal of Indo-European Studies 45(3/4): 267311.Google Scholar
Blažek, Václav, & Schwartz, Michal. 2016. The early Indo-Europeans in Central Asia and China: Cultural relations reflected in language. Innsbruck: IBK.Google Scholar
Boisacq, Emile. 1938. Dictionnaire étymologique de la langue grecque. Paris: Klincksieck.Google Scholar
Boroffka, Nikolaus. 1991. Die Verwendung von Eisen in Rumänien von den Anfängen bis in das 8. Jahrhundert v. Chr.: Vortrag gehalten zu Ehren von John Alexander auf dem Symposium ‘Europe in the 1st Millenium B.C.’, New York, 3rd4th April 1986, Institute of Archaeology Oxford. Berlin: self-published.Google Scholar
Boutkan, Dirk, & Kossmann, Maarten. 1999. Some Berber parallels of European substratum words. Journal of Indo-European Studies 27: 87100.Google Scholar
Boutkan, Dirk, & Kossmann, Maarten. 2001. On the etymology of ‘silver’. NOWELE 38: 315.CrossRefGoogle Scholar
Breyer, Gertraud. 1993. Etruskisches Sprachgut im Lateinischen unter Ausschluss des spezifisch onomastischen Bereiches. Leuven: Peeters.Google Scholar
Brixhe, Claude. 2004. Corpus des Inscriptions Paléo-Phrygiennes. Supplément II. Kadmos 43: 1130.Google Scholar
Brüch, Josef. 1914. Zwei ligurische Wörter im Lateinisch-Romanischen. Zeitschrift für vergleichende Sprachforschung auf dem Gebiete der indogermanischen Sprachen 46(4): 351373.Google Scholar
Brumlich, Markolf, Meyer, Michael, & Lychatz, Bernd. 2012. Archäologische und archäometallurgische Untersuchungen zur latènezeitlichen Eisenverhüttung im nördlichen Mitteleuropa. Praehistorische Zeitschrift 87(2): 433473.Google Scholar
Bugge, Sophus. 1893. Beiträge zur etymologischen erläuterung der armenischen sprache. Zeitschrift für vergleichende Sprachforschung 32(1): 187.Google Scholar
Burrow, T. 1972. A reconsideration of Fortunatov’s law. Bulletin of the School of Oriental and African Studies 34: 538559.Google Scholar
Buyaner, David B. On the etymology of the Old Iranian term for ‘iron’. In: Farridnejad, Shervin (ed.), zaraθuštrōtəma. Zoroastrian and Iranian studies in honour of Philip G. Kreyenbroek, 5167. Leiden: Brill.Google Scholar
Cantera, Alberto. 2017. The phonology of Iranian. In: Klein, Jared, Joseph, Brian, & Fritz, Matthias (ed.), Handbook of comparative and historical Indo-European linguistics, 481503. Berlin: de Gruyter Mouton.Google Scholar
Champion, Timothy. 2018. Iron and iron technology. In: Haselgrove, Colin, Rebay-Salisbury, Katharina, & Wells, Peter S (ed.), The Oxford handbook of the European Iron Age.Google Scholar
Chang, Kun. 1972. Sino-Tibetan ‘iron’ *qhleks. Journal of the American Oriental Society 92(3): 436446.CrossRefGoogle Scholar
Chernykh, E. N. 1992. Ancient metallurgy in the USSR: The early Metal Age. Cambridge: Cambridge University Press.Google Scholar
Chirikba, Viacheslav A. 1996. Common West Caucasian: The reconstruction of its phonological system and parts of its lexicon and morphology. Leiden: CNWS.Google Scholar
Civil, Miguel. 1976. Notes on Sumerian lexicography III. Journal of Cuneiform Studies 28(3): 183187.Google Scholar
Clauson, Gerard. 1972. An etymological dictionary of pre-thirteenth-century Turkish. Oxford: Clarendon Press.Google Scholar
Corretti, Alessandro, & Benvenuti, Marco. 2001. The beginning of iron metallurgy in Tuscany with special reference to Etruria Mineraria. Mediterranean Archaeology 14: 127145.Google Scholar
Cowgill, Warren C., & Mayrhofer, Manfred. 1986. Indo-Germanische Grammatik. Vol. 1.1. Halbband. Heidelberg: Winter.Google Scholar
Danti, Michael D. 2013. Late Bronze and Early Iron Age in northwestern Iran. In: Potts, D.T. (ed.), Oxford handbook of ancient Iran, 327376. Oxford: Oxford University Press.Google Scholar
Delamarre, Xavier. 2003. Dictionnaire de la langue gauloise. 2nd ed. Paris: Editions Errance.Google Scholar
DELG = Chantraine, Pierre. 1999. Dictionaire étymologique de la langue grecque: histoire des mots. 2nd ed. Paris: Klincksieck.Google Scholar
Derksen, Rick. 2007. Etymological dictionary of the Slavic inherited lexicon. Leiden & Boston: Brill.Google Scholar
Derksen, Rick. 2015. Etymological dictionary of the Baltic inherited lexicon. Leiden & Boston: Brill.Google Scholar
Deshayes, Albert. 2003. Dictionnaire étymologique du breton. Chasse-Marée: Douarnenez.Google Scholar
Diakonoff, Igor M. 1985. Hurro-Urartian borrowings in Old Armenian. Journal of the American Oriental Society 105(4): 597603.Google Scholar
Dossin, Georges. 1948. Le vocabulaire de Nuzi SMN 2559. Révue d’Assyriologie et d’archéologie orientale 42(1/2): 2134.Google Scholar
Dossin, Georges. 1971. Grèce et Orient. Révue belge de Philologie et d’Historie 49(1): 513.Google Scholar
Driessen, Michiel. 2003. *h2é-h2us-o-, the Proto-Indo-European term for ‘gold’. Journal of Indo-European Studies 31: 347362.Google Scholar
Dudarev, S. L. 2004. The mastering of iron-working by the peoples of the northern Caucasus in the Early Iron Age. Ancient West and East 3: 119.Google Scholar
DUL = Olmo Lete, Gregorio del & Sanmartín, Joaquín. 2015 . Dictionary of the Ugaritic language in the alphabetic tradition. 3rd ed. Translated and edited by Wilfred G.E. Watson. Leiden & Boston: Brill.Google Scholar
EDAIL = Martirosyan, Hrach. 2010. Etymological dictionary of the Armenian inherited lexicon. Leiden & Boston: Brill.Google Scholar
EDG = Beekes, Robert S. P. 2010. Etymological dictionary of Greek. Leiden & Boston: Brill.Google Scholar
EDHIL = Kloekhorst, Alwin. 2008. Etymological dictionary of the Hittite inherited lexicon. Leiden & Boston: Brill.Google Scholar
EDL = De Vaan, Michiel. 2008. Etymological dictionary of Latin and the other Italic languages. Leiden & Boston: Brill.Google Scholar
EDPC = Matasović, Ranko. 2009. Etymological dictionary of Proto-Celtic. Leiden & Boston: Brill.Google Scholar
EDPG = Kroonen, Guus. 2013. Etymological dictionary of Proto-Germanic. Leiden & Boston: Brill.Google Scholar
EIEC = Mallory, James P. & Adams, Douglas Q.. 1997. Encyclopedia of Indo-European culture. London & Chicago: Fitzroy Dearborn.Google Scholar
EM = Ernout, Alfred & Meillet, Antoine. 2001. Dictionnaire étymologique de la langue latine. 4th ed. Paris: Klincksieck.Google Scholar
Erb-Satullo, Nathaniel L. 2019. The innovation and adoption of iron in the ancient Near East. Journal of Archaeological Research 27: 557607.Google Scholar
Ernout, Alfred. 1946. Philologica. Vol. I. Paris: Klincksieck.Google Scholar
EWAia = Mayrhofer, Manfred. 1992–1996. Etymologisches Wörterbuch des Altindoarischen. 2 vols. Heidelberg: Winter.Google Scholar
Fick = Fick, August. 1890–1909. Vergleichendes Wörterbuch der indogermanischen Sprachen. 3 vols. Göttingen: Vandenhoeck & Rupprecht.Google Scholar
Flasdieck, Hermann M. 1952. Zinn und Zink: Studien zur abendländischen Wortgeschichte. Tübingen: Max Niemeyer Verlag.Google Scholar
Forbes, R. J. 1950. Metallurgy in Antiquity: A notebook for archaeologists and technologists. Leiden: Brill.Google Scholar
Fortunatov, Filipp F. 1881. L + dental im Altindischen. Beiträge zur Kunde der indogermanischen Sprachen 6: 215219.Google Scholar
Foxhall, Lin. 2018. The central Mediterranean and the Aegean. In: Haselgrove, Colin, Rebay-Salisbury, Katharina, & Wells, Peter S. (ed.), The Oxford handbook of the European Iron Age.CrossRefGoogle Scholar
Freeman, Philip. 1999. Homeric κασσίτερος. Glotta 75(3/4): 222225.Google Scholar
Furnée, Edzard J. 1972. Die wichtigsten konsonantischen Erscheinungen des Vorgriechischen. The Hague & Paris: De Gruyter Mouton.Google Scholar
Fortson, Benjamin W., IV. 2010. Indo-European language and culture: An introduction. 2nd ed. Malden (Mass).: Wiley-Blackwell.Google Scholar
Garcia, Dominique. 2018. Southern France. In: Haselgrove, Colin, Rebay-Salisbury, Katharina, & Wells, Peter S. (ed.), The Oxford handbook of the European Iron Age.Google Scholar
Garner, Jennifer. 2015. Bronze Age tin mines in Central Asia. In Hauptmann, Andreas & Modarressi-Tehrani, Diana (ed.), Archaeometallurgy in Europe III: Proceedings of the 3rd international conference, Deutsches Bergbau-Museum Bochum, June 29–July 1, 2011, 135143. Bochum: Vereinigung der Freunde von Kunst und Kultur im Bergbau.Google Scholar
Garnier, Romain. 2017. La dérivation inverse en latin. Innsbruck: IBS.Google Scholar
Gebauer, Anne Birgitte, Sørensen, Lasse, Taube, Michelle, & Wielandt, Daniel. 2020. First metallurgy in Northern Europe: An Early Neolithic crucible and a possible tuyère from Lønt, Denmark. European Journal of Archaeology: 1–21.Google Scholar
Georgiev, Vladimir. 1936. Lat. ferrum, griech. χαλκός, abg. želězo und Verwandtes. Zeitschrift für vergleichende Sprachforschung auf dem Gebiete der indogermanischen Sprachen 63(3/4): 250256.Google Scholar
Gerola, B. 1942. Substrato mediterraneo e latino. Studi Etruschi 16: 345368.Google Scholar
GEW = Frisk, Hjalmar. 1960–1972. Griechisches etymologisches Wörterbuch. 3 vols. Heidelberg: Winter.Google Scholar
Giardino, Claudio. 2005. Metallurgy in Italy between the Late Bronze Age and the Early Iron Age: The coming of iron. In: Attema, Peter, Nijboer, Albert, & Zifferero, Andrea (ed.), Papers in Italian archaeology VI: Communities and settlements from the Neolithic to the Early Medieval period, 491505. Oxford: Archaeopress.Google Scholar
Gimatzidis, Stefanos. 2018. Northern Greece and the central Balkans. In: Haselgrove, Colin, Rebay-Salisbury, Katharina, & Wells (, Peter S.ed.), The Oxford handbook of the European Iron Age.Google Scholar
Giusfredi, Federico. 2017. On Sumerian ku₃(.g)-an (a metal) and some allegedly derived words. Mémoires de NABU, March 2017: 13–15.Google Scholar
Goetze, Albrecht. 1947. Contributions to Hittite lexicography. Journal of Cuneiform Studies 1: 307320.Google Scholar
HAB = Ačaṙyan, Hračᶜeay. 1971–1976. Hayeren armatakan baṙaran. 4 vols. 2nd ed. Yerevan: Erevani Hamalsaran.Google Scholar
Häberl, Charles. n.d. A question of orthography: The Latino-Punic inscriptions. Unpublished. https://www.academia.edu/19367013/A_Question_of_Orthography_the_Latino-Punic_Inscriptions.Google Scholar
Häkkinen, Jaakko. 2009. Kantauralin ajoitus ja paikannus: perustelut puntarissa. Journal de la Société Finno-Ougrienne 92: 956.Google Scholar
Halleux, Robert. 1969. Lapis-lazuli, azurite ou pâte de verre? À propos de kuwano et kuwanowoko dans les tablettes mycéniennes. Studi micenei ed egeo-anatolici 9: 4563.Google Scholar
Harding, A. F. 2013. Trade and exchange. In: Fokkens, H. & Harding, A.F. (ed.), The Oxford handbook of the European Bronze Age, 370381. Oxford: Oxford University Press.Google Scholar
Haspelmath, Martin. 1993. A grammar of Lezgian. Berlin & New York: Mouton de Gruyter.CrossRefGoogle Scholar
HED = Puhvel, Jaan. 1984–. Hittite etymological dictionary. Berlin & New York: Mouton de Gruyter.Google Scholar
Hill, Eugen. 2003. Untersuchungen zum inneren Sandhi des Indogermanischen (Münchner Forschungen zur historischen Sprachwissenschaft 1). Bremen: Hempen Verlag.Google Scholar
Hinz, Walther, & Koch, Heidemarie. 1987. Elamisches Wörterbuch. Berlin: Reimer.Google Scholar
Hirt, Herman. 1905–1907. Die Indogermanen: Ihre Verbreitung, ihre Urheimat und ihre Kultur. 2 vols. Strassbourg: Trübner.Google Scholar
Hjärthner-Holdar, Eva. 1993. Järnets och järnmetallurgins introduktion i Sverige. Uppsala: Societas Archaeologica Upsaliensis.Google Scholar
Hoffmann, Karl, & Forssman, Bernhard. 1996. Avestische Laut- und Flexionslehre (Innsbrucker Beiträge zur Sprachwissenschaft 84). Innsbruck: Institut für Sprachwissenschaft.Google Scholar
Hoffner, Harry A. 1967. Review of Die Kaškäer: Ein Beitrag zur Ethnographie des alten Kleinasien by E. von Schuler (Berlin: de Gruyter, 1965). Journal of the American Oriental Society 87(2): 179185.Google Scholar
Hoffner, Harry A. 1968. A Hittite text in epic style about merchants. Journal of Cuneiform Studies 22(2): 3445.CrossRefGoogle Scholar
Holopainen, Sampsa. 2019. Indo-Iranian borrowings in Uralic: Critical overview of the sound substitutions and distribution criterion. PhD diss., University of Helsinki.Google Scholar
Hommel, Fritz. 1881. Allgemeine Zeitung 231 (August 19, 1881).Google Scholar
Hübschmann, Heinrich. 1897. Armenische Grammatik. 1. Teil: Armenische Etymologie. Leipzig: Breitkopf & Härtel.Google Scholar
Huld, Martin E. 2012. Some observations on the development of Indo-European metallurgy. In: Huld, M. E., Jones-Bley, K., & Miller, D. (ed.), Archaeology and language: Indo-European studies presented to James P. Mallory (JIES Monograph No. 60), 281356. Washington, DC: Institute for the Study of Man.Google Scholar
Hüsing, G. 1907. Miszellen. 4. Die Kassiteriden. Orientalistische Litteraturzeitung, January 1907, 1: 2526.Google Scholar
IEW = Pokorny, Julius. 1959. Indogermanisches etymologisches Wörterbuch. Bern: Francke.Google Scholar
Ivanov, Vjačeslav. 1977. O proisxoždenii nekotoryx baltijskix nazvanij metallov. Baltistica 13(1): 223236.Google Scholar
Iversen, Rune, & Kroonen, Guus 2017. Talking Neolithic: Linguistic and archaeological perspectives on how Indo-European was implemented in southern Scandinavia. American Journal of Archaeology 121(4): 511525.Google Scholar
Jagersma, Abraham H. 2010. A descriptive grammar of Sumerian. PhD diss., Leiden University.Google Scholar
J̌ahowkyan, Gevorg B. 1987. Hayocᶜ lezvi patmowtᶜyown: naxagrayin žamanakašrǰan. Yerevan: Erevani Hamalsaran.Google Scholar
Janhunen, Juha. 1983. On early Indo-European-Samoyed contacts. In: Janhunen, J., Peräniitty, A., & Suhonen, S. (ed.), Symposium saeculare Societas Fenno-Ugricae, 115127. Helsinki: Société Finno-Ougrienne.Google Scholar
Johannsen, Jens Winther. 2016. Heavy metal: Lead in Bronze Age Scandinavia. Fornvännen 111: 153161.Google Scholar
Kallio, Petri. 2004. Tocharian loanwords in Samoyed? In: Hyvärinen, Irma, Kallio, Petri, & Korhonen, Jarmo (ed.), Etymologie, Entlehnungen und Entwicklungen: Festschrift für Jorma Koivulehto zum 70. Geburtstag (Mémoires de la Société Néophilologique de Helsinki 63), 129137. Helsinki: Société Néophilologique.Google Scholar
Karsten, T. E. 1928. Die Germanen: Eine Einführung in die Geschichte ihrer Sprache und Kultur. Berlin & Leipzig: Walter de Gruyter & Co.Google Scholar
Kas’jan, A.C. 2010. Leksičeskie kontakty xattskogo jazyka. Indoevropejskoe jazykoznanie i klassičeskaja filologija 10(1): 445475.Google Scholar
Kauffmann, Friedrich. 1913. Deutsche Altertumskunde. 1. Hälfte: Von der Urzeit bis zur Völkerwanderung. Munich: Beck.Google Scholar
KEWA = Mayrhofer, Manfred. 1953. Kurzgefasstes etymologisches Wörterbuch des Altindischen. 4 vols. Heidelberg: Winter.Google Scholar
Klimov, Georgij A. 1964. Etimologičeskij slovar’ kartvel’skix jazykov. Moscow: Nauka.Google Scholar
Klingenschmitt, Gert. 2000. Mittelpersisch. In: Forssman, Bernhard & Plath, Robert (ed.), Indoarisch, Iranisch und die Indogermanistik: Arbeitstagung der Indogermanischen Gesellschaft vom 2 bis 5 Oktober 1997 in Erlangen, 191229. Wiesbaden: Reichert Verlag.Google Scholar
Klinger, Jörg. 1995. Hattisch. In: Streck, M. P. (ed.), Sprachen des alten Orients, 128134. Darmstadt: WBG.Google Scholar
Koch, John T. 2020. Celto-Germanic: Later prehistory and Post-Proto-Indo-European vocabulary in the north and west. Aberystwyth: Canolfan Uwchefrydiau Cymreig a Cheltaidd Prifysgol Cymru.Google Scholar
Korn, Agnes. 2003. Towards a historical grammar of Balochi: Studies in Balochi historical phonology and vocabulary. PhD diss., Goethe University Frankfurt.Google Scholar
Koryakova, Ludmila, Kuzminykh, Sergei & Beltikova, Galina. 2008. The introduction of iron technology into Central-Northern Eurasia. In: Forenius, Svante, Hjärthner-Holdar, Eva & Risberg, Christina (ed.), The introduction of iron in Eurasia: papers presented at the Uppsala Conference on October 4-8, 2001, 112127. Uppsala: Riksantikvarieämbetet.Google Scholar
Krogmann, Willy. 1937. Lat. ferrum. Zeitschrift für vergleichende Sprachforschung auf dem Gebiete der indogermanischen Sprachen 64(3/4): 267269.Google Scholar
Krogmann, Willy. 1940. Kelt. “omii̯o- „Erz, Kupfer”. Zeitschrift für Celtische Philologie 21: 4849.Google Scholar
Kümmel, Martin. 2017. Even more traces of the accent in Armenian? In: Hansen, Bjarne S.S. et al. (ed.), Usque ad Radices: Indo-European studies in honour of Birgit Anette Olsen, 439452. Copenhagen: Museum Tusculanum Press.Google Scholar
Kümmel, Martin. 2018. The survival of laryngeals in Iranian. In: van Beek, Lucien, Kloekhorst, Alwin, Kroonen, Guus, Peyrot, Michaël, & Pronk, Tijmen (ed.), Farnah: Indo-Iranian and Indo-European studies in honor of Sasha Lubotsky, 162172. Ann Arbor & New York: Beech Stave Press.Google Scholar
de Lamberterie, Charles. 1978. Armeniaca I–VIII: Études lexicales. Bulletin de la Société de Linguistique de Paris 73: 243283.Google Scholar
Lang, Valter. 2018. The eastern Baltic. In: Haselgrove, Colin, Rebay-Salisbury, Katharina, & Wells, Peter S. (ed.), The Oxford Handbook of the European Iron Age.Google Scholar
LEIA = Vendryes, Joseph, Bachallery, Édouard, & Lambert, Pierre-Yves. 1959–1996. Lexique étymologique de l’irlandais ancien. Dublin: Dublin Institute for Advanced Studies.Google Scholar
Levinsen, Karin. 1984. Jernets introduktion i Danmark. In: Kjærum, Poul (ed.), KUML 1982–1983: Årbog for Jysk Arkæologisk Selskab, 153168. Aarhus: Jysk Arkæologisk SelskabGoogle Scholar
LEW = Fraenkel, Ernst. 1962–1965. Litauisches etymologisches Wörterbuch. 2 vols. Heidelberg: Winter.Google Scholar
LIV² = Rix, Helmut, Kümmel, Martin, Zehnder, Thomas, Lipp, Reiner, & Schirmer, Brigitte (ed.). 2001. Lexikon der indogermanischen Verben. 2nd ed. Wiesbaden: Reichert.Google Scholar
Loma, Aleksandar. 2005. Zur Frage der frühesten griechisch-iranischen Sprachbeziehungen: Gr. κασσίτερος. In: Meiser, Gerhard & Hackstein, Olaf (eds.), Sprachkontakt und Sprachwandel: Akten der XI. Fachtagung der Indogermanischen Gesellschaft, 17.–23. September 2000, Halle an der Saale, 332340. Wiesbaden: Reichert.Google Scholar
LS = Lewis, Charlton T. & Short, Charles. 1897. A Latin dictionary. Oxford: Clarendon Press.Google Scholar
Lull, Vicente, Micó, Rafael, Herrada, Cristina Rihuete, & Risch, Roberto. 2014. The social value of silver in El Argar. In: Meller, H., Risch, R. & Pernicka, E. (ed.), Metalle der Macht: Frühes Gold und Silber: 6. Mitteldeutscher Archäologentag vom 17. bis 19. Oktober 2013 in Halle (Saale), 557576. Halle (Saale): Landesmuseum für Vorgeschichte.Google Scholar
Machek, Václav. 1957. Etymologický slovník jazyka českého a slovenského. Prague: Nakladatelství C̆eskoslovenské akademie věd.Google Scholar
Mallory, James P., & Huld, Martin E.. 1984. Proto-Indo-European ‘silver’. Zeitschrift für vergleichende Sprachforschung 97(1): 112.Google Scholar
McManus, Damian. 1991. A guide to Ogam. Maynooth: An Sagart.Google Scholar
Masson, Emilia. 1967. Recherches sur les plus anciens emprunts sémitiques en grec. Paris: Klincksieck.Google Scholar
Meillet, Antoine. 1923. Review of Baltisch-slawisches Wörterbuch by R. Trautmann (Göttingen: Vandenhoeck & Ruprecht, 1923). Bulletin de la Société de Linguistique 24(2): 135139.Google Scholar
Meillet, Antoine. 1936. Esquisse d’une grammaire comparée de l’arménien classique. 2nd ed. Vienna: Mekhitharistes.Google Scholar
Melchert, H. Craig. 2008. Greek mólybdos as a loanword from Lydian. In: Collins, B. J., Bachvarova, M. R., & Rutherford, I. C. (ed.), Anatolian interfaces: Hittites, Greeks and their neighbours, 153158. Oxford: Oxbow.Google Scholar
Metzner-Nebelsick, Carola. 2018. Central Europe. In: Haselgrove, Colin, Rebay-Salisbury, Katharina, & Wells, Peter S. (ed.), The Oxford Handbook of the European Iron Age.Google Scholar
Meyer-Lübke, Wilhelm. 1911. Romanisches etymologisches Wörterbuch. Heidelberg: Winter.Google Scholar
Militarev, Alexander, & Kogan, Leonid. 2000. Semitic etymological dictionary. Vol. I. Münster: Ugarit-Verlag.Google Scholar
Muller, F. 1918. Etymologiae Graecae. Mnemosyne 46(2): 135155.Google Scholar
Murtonen, A. 1989. Hebrew in its West Semitic setting. Leiden & New York: Brill.Google Scholar
NCED = Nikolayev, S. & Starostin, S. L.. 1994. A North Caucasian Etymological Dictionary. Moscow: Asterisk.Google Scholar
Neu, Erich. 1995. Zur Herkunft des Inselnames Kypros. Glotta 73(1): 17.Google Scholar
Nieling, Jens. 2009. Die Einführung der Eisentechnologie in Südkaukasien und Ostanatolien während der Spätbronze- und Früheisenzeit. Aarhus: Universitetsforlaget.Google Scholar
NIL = Wodtko, Dagmar S., Irslinger, Britta, & Schneider, Carolin. 2008. Nomina im indogermanischen Lexikon. Heidelberg: Winter.Google Scholar
Nørgaard, H. W., Pernicka, E., & Vandkilde, H.. 2019. On the trail of Scandinavia’s early metallurgy: Provenance, transfer and mixing. PLoS ONE 14(7): e0219574.Google Scholar
Nowakowski, Wojciech. 2018. Eastern Central Europe: Between the Elbe and the Dnieper. In: Haselgrove, Colin, Rebay-Salisbury, Katharina, & Wells, Peter S. (ed.), The Oxford Handbook of the European Iron Age. https://doi.org/10.1093/oxfordhb/9780199696826.013.10.Google Scholar
Olsen, Birgit A. 1999. The noun in biblical Armenian: origin and word-formation – with special emphasis on the Indo-European heritage. Berlin & New York: Mouton de Gruyter.Google Scholar
Orel, Vladimir. 1998. Albanian etymological dictionary. Leiden, Boston, & Cologne: Brill.Google Scholar
Papadopoulos, John K., Bejko, Lorenc, & Morris, Sarah P.. 2007. Excavations at the prehistoric tumulus of Lofkënd in Albania: A preliminary report from the 2004–2005 seasons. American Journal of Archaeology 11(1): 105147.Google Scholar
Pedersen, Holger. 1909. Vergleichende Grammatik der keltischen Sprachen. Bd. 1. Einleitung und Lautlehre. Göttingen: Vandenhoeck & Ruprecht.Google Scholar
Pedersen, Holger. 1924. Armenier. B. Sprache. In Ebert, Max (ed.), Reallexikon der Vorgeschichte, Vol. 1, 219226. Berlin: de Gruyter.Google Scholar
Petr, V. J. 1898. Über den Wechsel der Laute d und l im Lateinischen. Beiträge zur Kunde der indogermanischen Sprachen 25: 127158.Google Scholar
Peyrot, Michaël. 2018. Tocharian B etswe ‘mule’ and Eastern East Iranian. In van Beek, Lucien, Kloekhorst, Alwin, Kroonen, Guus, & Pronk, Tijmen (ed.), Farnah: Indo-Iranian and Indo-European studies in honor of Sasha Lubotsky, 270283. Ann Arbor and New York: Beech Stave Press.Google Scholar
Pfiffig, Ambros Josef. 1969. Die etruskische Sprache. Graz: Akademische Druck- und Verlagsanstalt.Google Scholar
Pigott, Vincent C. 1999. The archaeometallurgy of the Asian Old World. Philadelphia: University of Pennsylvania Museum.Google Scholar
Pinault, Georges-Jean. 2006. Further links between the Indo-Iranian substratum and BMAC. In: Tikkanen, Betil & Hettrich, Heinrich (ed.), Themes and tasks in Old Middle Indo-Aryan linguistics (Papers of the 12th World Sanskrit Conference, vol. 5), 167196. Delhi: Motilal Banarsidass Publishers.Google Scholar
Pisani, Vittore. 1959. Saggi di linguistica storica. Turin: Rosenberg & Sellier.Google Scholar
Pleiner, Radomír. 1996. Das frühe Eisen: Von den Kleinwaagemengen zu der ältesten Industrie. Ethnographisch-Archäologische Zeitschrift 37(3): 283291.Google Scholar
Pott, August Friedrich. 1833. Etymologische Forschungen auf dem Gebiete der indogermanischen Sprachen. Lemgo: Meyer.Google Scholar
Pronk, Tijmen. 2011. The “Saussure effect” in Indo-European languages other than Greek. Journal of Indo-European Studies 39: 176193.Google Scholar
Pronk, Tijmen, & Kloekhorst, Alwin. 2019. Introduction: Reconstructing Proto-Indo-Anatolian and Proto-Indo-Uralic. In: Pronk, T. & Kloekhorst, A. (ed.), The precursors of Proto-Indo-European: The Indo-Anatolian and Indo-Uralic hypotheses, 114. Leiden & Boston: Brill.Google Scholar
Räsänen, Martti. 1969. Versuch eines etymologischen Wörterbuchs der Türksprachen. Vol. 1. Helsinki: Suomalais-Ugrilainen Seura.Google Scholar
Remmer, Ulla. 2006. Frauennamen im Rigveda und im Avesta. Vienna: Verlag der Österreichischen Akademie der Wissenschaften.Google Scholar
REW = Vasmer, Max. 1953. Russisches etymologisches Wörterbuch. 2 vols. Heidelberg: Winter.Google Scholar
Richter, Thomas. 2012. Bibliographisches Glossar des Hurritischen. Wiesbaden: Harrassowitz.Google Scholar
Roberts, Benjamin W. 2009. Production networks and consumer choice in the earliest metal of Western Europe. Journal of World Prehistory 22: 461–81.Google Scholar
Rosół, Rafal. 2013. Frühe semitische Lehnwörter im Griechischen. Frankfurt am Main: Peter Lang.Google Scholar
Schaffner, Stefan. 2001. Das Vernersche Gesetz und der innerparadigmatische Wechsel des urgermanischen im Nominalbereich. Innsbruck: IBS.Google Scholar
Schaffner, Stefan. 2016/17. Lateinisch rutilus, rötlich, gelbrot, goldgelb’, altirisch ruithen, Strahl, Glanz’ und mittelkymrisch rwt ‘Rost, Korrosion.’ Die Sprache: Zeitschrift für Sprachwissenschaft 52(1): 101123.Google Scholar
Schrader, Otto. 1883. Sprachvergleichung und Urgeschichte: Linguistisch-historische Beiträge zur Erforschung des indogermanischen Altertums. Jena: Hermann Costenoble.Google Scholar
Schrijver, Peter. 1991. The reflexes of the Proto-Indo-European laryngeals in Latin. Amsterdam & Atlanta: Rodopi.Google Scholar
Schrijver, Peter. 2018. Talking Neolithic: The case for Hatto-Minoan and its relationship to Sumerian. In: Kroonen, G., Mallory, J. P., & Comrie, B. (ed.), Talking Neolithic: Proceedings of the workshop on Indo-European origins held at the Max Planck Institute for Evolutionary Anthropology, Leipzig, December 2–3, 2013 (JIES Monograph No. 65), 336374. Washington, DC: Institute for the Study of Man.Google Scholar
Schultze, Wolfgang. 2013. Historische und areale Aspekte der Bodenschatz-Terminologie in den ostkaukasischen Sprachen. Iran and the Caucasus 17: 295320.Google Scholar
Schuchardt, Hugo. 1913. Baskisch-hamitische Wortvergleichungen. Revista Internacional de Estudios Vascos/Revue International des Études Basques 7: 289340.Google Scholar
Segert, Stanislav. 1976. A grammar of Phoenician and Punic. Munich: C. H. Beck.Google Scholar
Seyer, Heinz. 1982. Siedlung und archäologische Kulture der Germanen im Havel-Spree-Gebiet in den Jahrhunderten vor Beginn u.Z. Berlin: Akademie Verlag.CrossRefGoogle Scholar
De Simone, Carlo. 1970. Die griechischen Entlehnungen im Etruskischen. Wiesbaden: Harrassowitz.Google Scholar
Soysal, Oǧuz. 2004. Hattischer Wortschatz in hethitischer Textüberlieferung. Leiden: Brill.Google Scholar
Soysal, Oǧuz. 2006. Das hethitische Wort für ‘Zinn’. Historische Sprachforschung 119: 109116.Google Scholar
Starostin, Sergei. 1985. Kul’turnaja leksika v obščeseverokavkazskom slovarnom fonde. In: Piotrovskij, B.B., Ivanov, V.V., & Ardzinba, V.G. (ed.), Drevnjaja Anatolija, 7494. Moscow: Nauka.Google Scholar
Stéphanidès, Michel. 1918. Petites contributions à l’histoire des sciences. Revue des Études Grecques 31(142): 197206.Google Scholar
Stifter, David. 1998. Study in red. Die Sprache: Zeitschrift für Sprachwissenschaft 40: 202223.Google Scholar
Stokes, Whitley & Bezzenberger, Adalbert. 1979. Wortschatz der keltischen Spracheinheit. 5th ed. Göttingen: Vandenhoeck & Ruprecht.Google Scholar
Szemerényi, Oswald. 1964. Syncope in Greek and Indo-European and the nature of the Indo-European accent. Naples: Instituto universitario orientale di Napoli.Google Scholar
Teržan, Biba, & de Marinis, Raffaele. 2018. The northern Adriatic. In: Haselgrove, Colin, Rebay-Salisbury, Katharina, & Wells, Peter S. (ed.), The Oxford Handbook of the European Iron Age. https://doi.org/10.1093/oxfordhb/9780199696826.013.10.Google Scholar
Tietze, Andreas. 2002. Tarihi ve Etimolojik Türkiye Türkçesi sözlüğü. Istanbul & Vienna: Simurg Kitapçılık, ÖAW.Google Scholar
Tomaschek, Wilhelm. 1884. Review of Schrader 1883. Litteratur-Blatt für orientalische Philologie 1 (Oct. 1883–Sept. 1884): 121–30.Google Scholar
Trask, R. L. 2008. Etymological dictionary of Basque. Edited for web publication by Max W. Wheeler. University of Sussex.Google Scholar
Trautmann, Reinhold. 1910. Die altpreussischen Sprachdenkmäler. Göttingen: Vandenhoeck und Ruprecht.Google Scholar
Tremblay, Xavier. 2004. Chalcographie: Sur χαλκός, lit geležìs et turc qoruγžin. Historische Sprachforschung 117(2): 238248.Google Scholar
Tremblay, Xavier. 2005. Irano-Tocharica et Tocharo-Iranica. Bulletin of SOAS 68(3): 421449.Google Scholar
Tripathi, D. N. 1996. Tin in the ancient world: a literary study. In: Singh, Sarva Daman (ed.), Culture through the ages: Prof. B.N. Puri felicitation volume, 161167. Delhi: Agam Kala Prakashan.Google Scholar
Trubachev, Oleg N. 1967. Iz slavjano-iranskix leksičeskix otnošenij. Ètimologija 1965: 381.Google Scholar
UEW = Rédei, Károly. 1986–1991. Uralisches etymologisches Wörterbuch. 3 vols. Budapest: Akadémiai Kiadó.Google Scholar
Valério, M., and Yakubovich, I.. 2010. Semitic word for ‘iron’ as Anatolian loanword. In: Nikolaev, T. M. (ed.), Isseledovanija po lingvistike i semiotike: Sbornik statej k jubileju Vyač. Vs. Ivanova, 108116. Moscow: Jazyki slavjanskix kul’tur.Google Scholar
Vaniček, Alois. 1881. Etymologisches Wörterbuch der lateinischen Sprache. Leipzig: Teubner.Google Scholar
Vanséveren, Sylvie. 2012. Noms de métaux dans les textes hittites. Anatolica 38: 203219.Google Scholar
Viitso, Tiit-Rein. 2013. Early metallurgy in language: The history of metal names in Finnic. In: Grünthal, Riho & Kallio, Petri (ed.), A linguistic map of prehistoric Northern Europe, 185200. Helsinki: Suomalais-Ugrilainen Seura.Google Scholar
WH = Walde, Alois & Hofmann, Johann Baptist. 1965–1972. Lateinisches Etymologische Wörterbuch. Heidelberg: Winter.Google Scholar
Wegner, Ilse. 2000. Einführung in die hurritische Sprache. Wiesbaden: Harrassowitz.Google Scholar
Weiss, Michael. 2020. Outline of the historical and comparative grammar of Latin. 2nd ed. Ann Arbor & New York: Beech Stave Press.Google Scholar
Witczak, Krzysztof Tomasz. 2009. A wandering word for ‘hardened iron, steel’: A study in the history of concepts and words. Studia Etymologica Cracoviensia 14: 291305.Google Scholar
van Windekens, Albert J. 1958. Pelasgisch und Westgermanisch: Neues Material. Die Sprache 4: 128138.Google Scholar
Yalçın, Ünsal. 1999. Early iron metallurgy in Anatolia. Anatolian Studies 49: 177187.Google Scholar
Zapatero, Gonzalo Ruiz, Manuel Fernández-Götz, , & Álvarez-Sanchís, Jesús. 2012. Die Ausbreitung der Eisenmetallurgie auf der Iberischen Halbinsel. In: Kern, Anton et al. (ed.), Technologieentwicklung und -transfer in der Hallstatt- und Latènezeit, 149166. Lagenweissbach: Beier & Beran.Google Scholar
Zhivlov, Mikhail. 2014. Studies in Uralic vocalism III. Journal of Language Relationship 12: 113148.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
×