Hostname: page-component-848d4c4894-pftt2 Total loading time: 0 Render date: 2024-06-01T23:47:21.641Z Has data issue: false hasContentIssue false

The Road Not Taken: How Early Landscape Learning and Adoption of a Risk-Averse Strategy Influenced Paleoindian Travel Route Decision Making in the Upper Ohio Valley

Published online by Cambridge University Press:  30 December 2020

Matthew P. Purtill*
Affiliation:
Department of Geology and Environmental Science, SUNY-Fredonia, 280 Central Avenue, Fredonia, NY 14063, USA
*
(matthew.purtill@fredonia.edu, corresponding author)

Abstract

To evaluate a model of the travel-route selection process for upper Ohio Valley Paleoindian foragers (13,500–11,400 cal BP), this study investigates archaeological data through the theoretical framework of landscape learning and risk-sensitive analysis. Following initial trail placement adjacent to a highly visible escarpment landform, Paleoindians adopted a risk-averse strategy to minimize travel outcome variability when wayfaring between Sandy Springs, a significant Ohio River Paleoindian site, and Upper Mercer–Vanport chert quarries of east-central Ohio. Although a least-cost analysis indicates an optimal route through the lower Scioto Valley, archaeological evidence for this path is lacking. Geomorphic and archaeological data further suggest that site absence in the lower Scioto Valley is not entirely due to sampling bias. Instead, evidence indicates that Paleoindians preferred travel within the Ohio Brush Creek–Baker's Fork valley despite its longer path distance through more rugged, constricted terrain. Potential travel through the lower Scioto Valley hypothesizes high outcome variability due to the stochastic nature of the late Pleistocene hydroregime. In this case, perceived outcome variability appears more influential in determining travel-route decisions among Paleoindians than direct efforts to reduce energy and time allocation.

Para evaluar un modelo del proceso de selección de ruta de tránsito para los recolectores paleoindios del valle superior del Ohio (13.500-11.400 cal BP), este estudio investiga datos arqueológicos a través del marco teórico de aprendizaje del paisaje y análisis del riesgo posible. Siguiendo una ubicación inicial de la senda adyacente a una formación escarpada muy visible, los paleo-indios adoptaron una estrategia adversa al riesgo para minimizar la variabilidad del resultado de tránsito en sus expediciones entre Sandy Springs, un importante yacimiento paleoindio en el río Ohio, y las excavaciones de sílex de Mercer-Vanport superior en el centro-este de Ohio. Aunque un análisis del menor coste indica una ruta óptima a través del valle bajo del Scioto, falta evidencia arqueológica de ese recorrido. Los datos geomórficos y arqueológicos sugieren además que la ausencia de yacimientos en el valle bajo del Scioto no se debe completamente al sesgo del muestreo. En su lugar, la evidencia demuestra que los paleoindios preferían desplazarse por el interior del valle Ohio Brush Creek-Baker's Fork a pesar de ser una distancia mayor a través de terreno más escabroso y restringido. Un recorrido potencial a través del valle bajo del Scioto proponía una alta variabilidad de resultados debido a la naturaleza aleatoria del régimen hídrico en el bajo Pleistoceno. En este caso, la percepción de variabilidad de resultado era más influyente en la elección de la ruta de tránsito entre los paleoindios que los esfuerzos directos de reducir la asignación de energía y tiempo.

Type
Articles
Copyright
Copyright © The Author(s), 2020. Published by Cambridge University Press on behalf of the Society for American Archaeology

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 Cited

Anderson, David G. 1990 The Paleoindian Colonization of Eastern North America. Research in Economic Anthropology Supplement 5:163216.Google Scholar
Anderson, David G. 1995 Paleoindian Interaction Networks in the Eastern Woodlands. In Native American Interaction: Multiscalar Analyses and Interpretations in the Eastern Woodlands, edited by Nassaney, Michael S. and Sassaman, Kenneth E., pp. 126. University of Tennessee Press, Knoxville.Google Scholar
Anderson, David G. 1996 Models of Paleoindian and Early Archaic Settlement in the Lower Southeast. In The Paleoindian and Early Archaic Southeast, edited by Anderson, David G. and Sassaman, Kenneth E., pp. 2957. University of Alabama Press, Tuscaloosa.Google Scholar
Anderson, David G. 2012 Least Cost Pathway Analysis in Archaeological Research: Approaches and Utility. In Least Cost Analysis of Social Landscapes: Archaeological Case Studies, edited by White, Devin Alan and Surface-Evans, Sarah, pp. 239258. University of Utah Press, Salt Lake City.Google Scholar
Anderson, David G., Bissett, Thaddeus G., and Yerka, Stephen J. 2014 The Late-Pleistocene Human Settlement of Interior North America: The Role of Physiography and Sea-Level Change. In Paleoamerican Odyssey, edited by Graf, Kelly E., Ketron, Caroline V., and Waters, Michael R., pp. 183203. Texas A&M University Press, College Station.Google Scholar
Anderson, David G., and Faught, Michael K. 1998 The Distribution of Fluted Paleoindian Projectile Points: Update 1998. Archaeology of Eastern North America 26:163187.Google Scholar
Anderson, David G., and Faught, Michael K. 2000 Palaeoindian Artifact Distributions: Evidence and Implications. Antiquity 74:507513.CrossRefGoogle Scholar
Anderson, David G., and Gillam, J. Christopher 2000 Paleoindian Colonization of the Americas: Implications from an Examination of Physiography, Demography, and Artifact Distribution. American Antiquity 65:4366.CrossRefGoogle Scholar
Anderson, David G., Smallwood, Ashley M., and Miller, D. Shane 2015 Pleistocene Human Settlement in the Southeastern United States: Current Evidence and Future Directions. PaleoAmerica 1:751.CrossRefGoogle Scholar
Anthony, David W. 1990 Migration in Archeology: The Baby and the Bathwater. American Anthropologist 92:895914.CrossRefGoogle Scholar
Anthony, David W. 1997 Prehistoric Migration as Social Process. In Migrations and Invasions in Archaeological Explanation, edited by Chapman, John and Hamerow, Helena, pp. 2132. BAR International Series 664. Archaeopress, Oxford.Google Scholar
Arbogast, Alan F., Bookout, Juleigh R., Schrotenboer, Bradley R., Lansdale, Amy, Rust, Ginny L., and Bato, Victorino A. 2008 Post-Glacial Fluvial Response and Landform Development in the Upper Muskegon River Valley in North-Central Lower Michigan, U.S.A. Geomorphology 102:615623.CrossRefGoogle Scholar
Armantier, Olivier 2006 Estimates of Own Lethal Risks and Anchoring Effects. Journal of Risk and Uncertainty 32:3756.CrossRefGoogle Scholar
Ashmore, Peter E. 1991 How Do Gravel-Bed Rivers Braid? Canadian Journal of Earth Sciences 28:326341.CrossRefGoogle Scholar
Bluhm, Lara E., and Surovell, Todd A. 2019 Validation of a Global Model of Taphonomic Bias Using Geologic Radiocarbon Ages. Quaternary Research 91:325328.CrossRefGoogle Scholar
Boulanger, Matthew T., Buchanan, Briggs, O'Brien, Michael J., Redmond, Brian G., Glascock, Michael D., and Eren, Metin I. 2015 Neutron Activation Analysis of 12,900-Year-Old Stone Artifacts Confirms 450–510 km Clovis Tool-Stone Acquisition at Paleo Crossing (33ME274), Northeast Ohio, USA. Journal of Archaeological Science 53:550558.CrossRefGoogle Scholar
Braun, E. Lucy 1928 Glacial and Post-Glacial Plant Migrations Indicated by Relic Colonies of Southern Ohio. Ecology 9:284302.CrossRefGoogle Scholar
Brockman, C. Scott 2006 Physiographic Regions of Ohio. State of Ohio, Department of Natural Resources, Division of Geological Survey, Columbus.Google Scholar
Broster, John B., Norton, Mark R., Shane Miller, D., Tune, Jesse W., and Baker, Jon D. 2013 Tennessee's Paleoindian Record: The Cumberland and Lower Tennessee River Watersheds. In In the Eastern Fluted Point Tradition, edited by Gingerich, Joseph A. M., pp. 299314. University of Utah Press, Salt Lake City.Google Scholar
Bruno, David, Lamb, Lisa, and Kaiwari, Jack 2014 The Landscapes of Mobility: The Flow of Place. In The Oxford Handbook of the Archaeology and Anthropology of Hunter-Gatherers, edited by Cummings, Vicki, Jordan, Peter, and Zvelebil, Marek, pp. 11631190. Oxford University Press, Oxford.Google Scholar
Buchanan, Briggs 2003 The Effects of Sample Bias on Paleoindian Fluted Point Recovery in the United States. North American Archaeologist 24:311338.CrossRefGoogle Scholar
Burmeister, Stefan 2000 Archaeology and Migration: Approaches to an Archaeological Proof of Migration. Current Anthropology 41:539567.CrossRefGoogle Scholar
Cabana, Graciela S., and Clark, Jeffery J. (editors) 2011 Rethinking Anthropological Perspectives on Migration. University Press of Florida, Gainesville.CrossRefGoogle Scholar
Carlson, Ernest H. 1991 Minerals of Ohio. Bulletin 69. State of Ohio, Department of Natural Resources, Division of Geological Survey, Columbus.Google Scholar
Carr, Dillon H., and Boszhardt, Robert F. 2010 Silver Mound, Wisconsin: Source of Hixton Silicified Sandstone. Midcontinental Journal of Archaeology 35:536.CrossRefGoogle Scholar
Cunningham, Roger M. 1973 Paleo-Hunters along the Ohio River. Archaeology of Eastern North America 1:118126.Google Scholar
Dyke, Arthur S., Moore, Andrew, and Robertson, Louis 2003 Deglaciation of North America. Geological Survey of Canada, Open File 1574, Ottawa, Ontario.CrossRefGoogle Scholar
Ellis, Christopher J. 2011 Measuring Paleoindian Range Mobility and Land-Use in the Great Lakes/Northeast. Journal of Anthropological Archaeology 30:385401.CrossRefGoogle Scholar
Ellis, Christopher J., Carr, Dillon H., and Loebel, Thomas J. 2011 The Younger Dryas and Late Pleistocene Peoples of the Great Lakes Region. Quaternary International 242:534545.CrossRefGoogle Scholar
Erber, Nathan R., Spahr, Paul N., and Martin, Dean R. 2016 Surficial Geology of the Chillicothe East Quadrangle, Ohio. Digital Map Series SG-4A. Ohio Division of Geological Survey, Columbus.CrossRefGoogle Scholar
Eren, Metin I., Buchanan, Briggs, and O'Brien, Michael J. 2015 Social Learning and Technological Evolution during the Clovis Colonization of the New World. Journal of Human Evolution 80:159170.CrossRefGoogle ScholarPubMed
Eren, Metin I., Chao, Anne, Chiu, Chun-Huo, Colwell, Robert K., Buchanan, Briggs, Boulanger, Matthew T., Darwent, John, and O'Brien, Michael J. 2016 Statistical Analysis of Paradigmatic Class Richness Supports Greater Paleoindian Projectile-Point Diversity in the Southeast. American Antiquity 81:174192.CrossRefGoogle Scholar
Eren, Metin I., Miller, G. Logan, Buchanan, Briggs, Boulanger, Matthew T., Bebber, Michelle R., Redmond, Brian G., Stephens, Charles “Chuck”, Coates, Lisa, Boser, Patricia, Sponseller, Becky, and Slicker, Matt 2019 The Black Diamond Site, Northeast Ohio, USA: A New Clovis Occupation in a Proposed Secondary Staging Area. Journal of Paleolithic Archaeology 2:211233.CrossRefGoogle Scholar
Fenneman, Nevin M. 1928 Physiographic Divisions of the United States. Annals of the Association of American Geographers 18:264354.CrossRefGoogle Scholar
Field, Andy 2009 Discovering Statistics Using SPSS. Sage, Los Angeles.Google Scholar
Gingerich, Joseph A. M., and Kitchel, Nathaniel R. 2015 Early Paleoindian Subsistence Strategies in Eastern North America: A Continuation of the Clovis Tradition? Or Evidence of Regional Adaptations? In Clovis: On the Edge of a New Understanding, edited by Smallwood, Ashley M. and Jennings, Thomas A., pp. 297318. Texas A&M University Press, College Station.Google Scholar
Golledge, Reginald G. 2003 Human Wayfinding and Cognitive Maps. In Colonization of Unfamiliar Landscapes: The Archaeology of Adaptation, edited by Rockman, Marcy and Steele, James, pp. 2543. Routledge, London.Google Scholar
Goodyear, Albert C. 1989 A Hypothesis for the Use of Cryptocrystalline Raw Materials among Paleoindian Groups of North America. In Eastern Paleoindian Lithic Resource Use, edited by Ellis, Christopher and Lothrop, Jonathan, pp. 110. Westview Press, Boulder, Colorado.Google Scholar
Güimil-Fariña, Alejadro, and Parcero-Oubiña, César 2015 “Dotting the Joins”: A Non-reconstructive Use of Least Cost Paths to Approach Ancient Roads: The Case of the Roman Roads in the NW Iberian Peninsula. Journal of Archaeological Science 54:3144.CrossRefGoogle Scholar
Gustas, Robert, and Supernant, Kisha 2019 Coastal Migration into the Americas and Least Cost Path Analysis. Journal of Anthropological Archaeology 54:192206.CrossRefGoogle Scholar
Herrmann, Edward W. 2013 The Geoarchaeology of Paleoindian and Early Archaic Sites in the White River Valley, Indiana. PhD dissertation, Department of Anthropology, Indiana University, Bloomington.Google Scholar
Herrmann, Edward W., and Monaghan, G. William 2019 Post-Glacial Drainage Basin Evolution in the Midcontinent, North America: Implications for Prehistoric Human Settlement Patterns. Quaternary International 511:6877.CrossRefGoogle Scholar
Herzog, Irmela 2013 The Potential and Limits of Optimal Path Analysis. In Computational Approaches to Archaeological Spaces, edited by Bevan, Andrew and Lake, Mark, pp. 179211. Left Coast Press, Walnut Creek, California.Google Scholar
Holliday, Vance T., and Miller, D. Shane 2014 The Clovis Landscape. In Paleoamerican Odyssey, edited by Graf, Kelly E., Ketron, Caroline V., and Waters, Michael R., pp. 221245. Texas A&M University Press, College Station.Google Scholar
Jennings, Thomas A. 2015 Clovis Adaptations in the Great Plains. In Clovis: On the Edge of a New Understanding, edited by Smallwood, Ashley M. and Jennings, Thomas A., pp. 277296. Texas A&M University Press, College Station.Google Scholar
Johnson, Eileen, Politis, Gustavo, Gutíerrrez, Maria, Martínez, Gustavo, and Miotti, Laura 2006 Grassland Archaeology in the Americas: From the U.S. Southern Plains to the Argentinean Pampas. In Paleoindian Archaeology: A Hemispheric Perspective, edited by Morrow, Juliet E. and Gnecco, Cristóbal, pp. 4468. University Press of Florida, Gainesville.Google Scholar
Kahneman, Daniel, and Tversky, Amos 1979 Prospect Theory: An Analysis of Decision under Risk. Econometrica 47:263292.CrossRefGoogle Scholar
Karanasiou, Angeliki, Moreno, Natalia, Moreno, Teresa, Viana, M., de Leeuw, Frank, and Querol, Xavier 2012 Health Effects from Sahara Dust Episodes in Europe: Literature Review and Research Gaps. Environment International 47:107114.CrossRefGoogle ScholarPubMed
Kelly, Robert L. 2003 Colonization of New Land by Hunter-Gatherers. In Colonization of Unfamiliar Landscapes: The Archaeology of Adaptation, edited by Rockman, Marcy and Steele, James, pp. 4458. Routledge, London.Google Scholar
Kelly, Robert L. 2013 The Lifeways of Hunter-Gatherers: The Foraging Spectrum. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Kelly, Robert L., and Todd, Lawrence C. 1988 Coming into the Country: Early Paleoindian Hunting and Mobility. American Antiquity 53:231244.CrossRefGoogle Scholar
Kempton, John P., and Goldthwait, Richard P. 1959 Glacial Outwash Terraces of the Hocking and Scioto River Valleys, Ohio. The Ohio Journal of Science 59:135151.Google Scholar
Knox, James C. 1995 Fluvial Systems since 20,000 yrs BP. In Global Continental Palaeohydrology, edited by Gregory, Kenneth J., Starkel, Leszek, and Baker, Vincent R., pp. 87108. Wiley, New York.Google Scholar
Krist, Frank J., and Brown, Daniel G. 1994 GIS Modeling of Paleo-Indian Period Caribou Migrations and Viewsheds in Northeastern Lower Michigan. Photogrammetric Engineering and Remote Sensing 60:11291137.Google Scholar
Lane, Leon, and Anderson, David G. 2001 Paleoindian Occupations of the Southern Appalachians: A View from the Cumberland Plateau of Kentucky and Tennessee. In Archaeology of the Appalachian Highlands, edited by Sullivan, Lynne P. and Prezzano, Susan C., pp. 88102. University of Tennessee Press, Knoxville.Google Scholar
Lepper, Bradley T. 2005 Pleistocene Peoples of Midcontinental North America. In Ice Age Peoples of North America: Environments, Origins, and Adaptations of the First Americans, edited by Bonnischesen, Robson and Turnmire, Karen, pp. 362394. Center for the Study of the First Americans, Oregon State University, Corvallis.Google Scholar
Livingood, Patrick 2012 No Crows Made Mounds: Do Cost-Distance Calculations of Travel Time Improve Our Understanding of Southern Appalachian Polity Size? In Least Cost Analysis of Social Landscapes: Archaeological Case Studies, edited by White, Devin Alan and Surface-Evans, Sarah, pp. 174187. University of Utah Press, Salt Lake City.Google Scholar
Loebel, Thomas J. 2012 Pattern or Bias? A Critical Evaluation of Midwestern Fluted Point Distributions Using Raster Based GIS. Journal of Archaeological Science 39:12051217.CrossRefGoogle Scholar
Lothrop, Jonathan C., Burke, Adrian L., Winchell-Sweeney, Susan, and Gauthier, Gilles 2018 Coupling Lithic Sourcing with Least Cost Path Analysis to Model Paleoindian Pathways in Northeastern North America. American Antiquity 83:462484.CrossRefGoogle Scholar
Loyola, Rodrigo, Núñez, Lautaro, and Cartajena, Isabel 2019 What's It Like Out There? Landscape Learning during the Early Peopling of the Highlands of the South-Central Atacama Desert. Quaternary International 533:724.CrossRefGoogle Scholar
Malard, Florian, Uehlinger, Urs, Zah, Rainer, and Tockner, Klement 2006 Flood-Pulse and Riverscape Dynamics in a Braided Glacial River. Ecology 87:704716.CrossRefGoogle Scholar
Meltzer, David J. 1985 On Stone Procurement and Settlement Mobility in Eastern Fluted Point Groups. North American Archaeologist 6:124.CrossRefGoogle Scholar
Meltzer, David J. 2002 What Do You Do When No One's Been There Before? Thoughts on the Exploration and Colonization of New Lands. In The First Americans: The Pleistocene Colonization of the New World, edited by Jablonski, Nina G., pp. 2758. California Academy of Sciences, San Francisco.Google Scholar
Meltzer, David J. 2003 Lessons in Landscape Learning. In Colonization of Unfamiliar Landscapes: The Archaeology of Adaptation, edited by Rockman, Marcy and Steele, James, pp. 222241. Routledge, London.Google Scholar
Meltzer, David J. 2004 Modeling the Initial Colonization of the Americas: Issues of Scale, Demography, and Landscape Learning. In Settlement of the American Continents: A Multidisciplinary Approach to Human Biogeography, edited by Barton, C. Michael, Clark, Geoffrey A., Yesner, David R., and Pearson, Georges A., pp. 123137. University of Arizona Press, Tuscon.Google Scholar
Meltzer, David J. 2009 First Peoples in a New World: Colonizing Ice Age America. University of California Press, Oakland.CrossRefGoogle Scholar
Meltzer, David J., and Holliday, Vance T. 2010 Would North American Paleoindians Have Noticed Younger Dryas Age Climate Changes? Journal of World Prehistory 23:141.CrossRefGoogle Scholar
Miller, D. Shane 2016 Modeling Clovis Landscape Use and Recovery Bias in the Southeastern United States Using the Paleoindian Database of the Americas (PIDBA). American Antiquity 81:697716.CrossRefGoogle Scholar
Miller, G. Logan, Bebber, Michelle R., Rutkoski, Ashley, Haythorn, Richard, Boulanger, Matthew T., Buchanan, Briggs, Bush, Jennifer, Owen Lovejoy, C., and Eren, Metin I. 2019 Hunter-Gatherer Gatherings: Stone-Tool Microwear from the Welling Site (33-Co-2), Ohio, U.S.A. Supports Clovis Use of Outcrop-Related Base Camps during the Pleistocene Peopling of the Americas. World Archaeology 51:4775.CrossRefGoogle Scholar
Miller, D. Shane, and Carmody, Stephen B. 2016 Colonization after Clovis: Using the Ideal Free Distribution to Interpret the Distribution of Late Pleistocene and Early Holocene Archaeological Sites in the Duck River Valley, Tennessee. Tennessee Archaeology 8:78101.Google Scholar
Mills, William C. 1914 Archeological Atlas of Ohio. Ohio State Archeological and Historical Society, Columbus.Google Scholar
Mitsakou, Christina, Kallos, George, Papantoniou, N., Spyrou, C., Solomos, Stavos, Astitha, Marina, and Housiadas, C. 2008 Saharan Dust Levels in Greece and Received Inhalation Doses. Atmospheric Chemistry and Physics 8:71817192.CrossRefGoogle Scholar
Morgan, Christopher 2015 Is It Intensification Yet? Current Archaeological Perspectives on the Evolution of Hunter-Gatherer Economies. Journal of Archaeological Research 23:163213.CrossRefGoogle Scholar
Mukherjee, Sumitava, Sahay, Arvind, Chandrasekhar Pammi, V. S., and Srinivasan, Narayanan 2017 Is Loss-Aversion Magnitude-Dependent? Measuring Prospective Affective Judgments Regarding Gains and Losses. Judgment and Decision Making 12:8189.CrossRefGoogle Scholar
Osorio, Daniela, Steele, James, Sepulveda, Marcela, Gayo, Eugenia, Capriles, Jose M., Katherine, Herrera, Ugalde, Paula, De Pol-Holz, Ricardo, Latorre, Claudio, and Santoro, Calogero M. 2017 The Dry Puna as an Ecological Megapatch and the Peopling of South America: Technology, Mobility, and the Development of a Late Pleistocene/Early Holocene Andean Hunter-Gatherer Tradition in Northern Chile. Quaternary International 461:4153.CrossRefGoogle Scholar
Parish, Ryan M. 2018 Lithic Procurement Patterning as a Proxy for Identifying Late Paleoindian Group Mobility along the Lower Tennessee River Valley. Journal of Archaeological Science: Reports 22:313323.Google Scholar
Pavey, Richard R., Goldthwait, Richard P., Scott Brockman, C., Hull, Dennis N., Mac Swinford, E., and Van Horn, Robert G. 1999 Quaternary Geology of Ohio. Map No. 2. State of Ohio, Department of Natural Resources, Division of Geological Survey, Columbus.Google Scholar
Portsmouth Public Library 2020 1940 Portsmouth Flood-Scioto River. Local History Digital Collection. Electronic document, https://www.yourppl.org/history/items/show/4365, accessed June 4, 2020.Google Scholar
Prasciunas, Mary M. 2011 Mapping Clovis: Projectile Points, Behavior, and Bias. American Antiquity 76:107126.CrossRefGoogle Scholar
Prufer, Olaf H., and Baby, Raymond S. 1963 The Paleo-Indians of Ohio. Ohio Historical Society, Columbus.Google Scholar
Prufer, Olaf H., and Wright, Norman L. 1970 The Welling Site (33co-2): A Fluted Point Workshop in Coshocton County, Ohio. Ohio Archaeologist 20:259268.Google Scholar
Purtill, Matthew P. 2009 The Ohio Archaic: A Review. In Archaic Societies: Diversity and Diversity across the Midcontinent, edited by Emerson, Thomas E., McElrath, Dale T., and Fortier, Andrew C., pp. 565606. SUNY Press, Albany, New York.Google Scholar
Purtill, Matthew P. 2017 Reconsidering the Potential Role of Saline Springs in the Paleoindian Occupation of Sandy Springs, Adams County, Ohio. Journal of Archaeological Science: Reports 13:164174.Google Scholar
Purtill, Matthew P., Steven Kite, J., and Forman, Steven 2019 Geochronology and Depositional History of the Sandy Springs Aeolian Landscape in the Unglaciated Upper Ohio River Valley, United States. Frontiers in Earth Science 7:322.CrossRefGoogle Scholar
Rademaker, Kurt, Reid, David A., and Bromley, Gordon R. M. 2012 Connecting the Dots: Least Cost Analysis, Paleogeography, and the Search for Paleoindian Sites in Southern Highland Peru. In Least Cost Analysis of Social Landscapes: Archaeological Case Studies, edited by White, Devin Alan and Surface-Evans, Sarah, pp. 3245. University of Utah Press, Salt Lake City.Google Scholar
Redmond, Brian G., and Tankersley, Kenneth B. 2005 Evidence of Early Paleoindian Bone Modification and Use at the Sheriden Cave Site (33WY252), Wyandot County, Ohio. American Antiquity 70:503526.CrossRefGoogle Scholar
Riley, Shawn J., DeGloria, Stephen D., and Elliot, Robert 1999 A Terrain Ruggedness Index That Quantifies Topographic Heterogeneity. Intermountain Journal of Sciences 5(1–4):2327.Google Scholar
Rockman, Marcy 2003 Knowledge and Learning in the Archaeology of Colonization. In The Colonization of Unfamiliar Landscapes: The Archaeology of Adaptation, edited by Rockman, Marcy and Steele, James, pp. 324. Routledge, New York.CrossRefGoogle Scholar
Rockman, Marcy 2009 Landscape Learning in Relation to Evolutionary Theory. In Macroevolution in Human Prehistory: Evolutionary Theory and Processual Archaeology, pp. Prentiss, Anna Marie, Kuijt, Ian, and Chatters, James C., pp. 5171. Springfield, New YorkCrossRefGoogle Scholar
Rode, Catrin, Cosmides, Leda, Hell, Wolfgang, and Tooby, John 1999 When and Why Do People Avoid Unknown Probabilities in Decisions under Uncertainty? Testing Some Predictions from Optimal Foraging Theory. Cognition 72:269304.CrossRefGoogle ScholarPubMed
Rosengreen, Theodore E. 1974 Glacial Geology of Highland County, Ohio. Ohio Division of Geological Survey, Columbus.Google Scholar
Rutledge, E. M., Holowychuk, N., Hall, G. F., and Wilding, L. P. 1975 Loess in Ohio in Relation to Several Possible Source Areas: I. Physical and Chemical Properties. Soil Science Society of America Journal 39:11251132.CrossRefGoogle Scholar
Seeman, Mark F., and Prufer, Olaf H. 1982 An Updated Distribution of Ohio Fluted Points. Midcontinental Journal of Archaeology 7:155169.Google Scholar
Seeman, Mark F., Summers, Gary, Dowd, Elaine, and Morris, Larry 1994 Fluted Point Characteristics at Three Large Sites: The Implications for Modeling Early Paleoindian Settlement Patterns in Ohio. In The First Discovery of America: Archaeological Evidence of the Early Inhabitants of the Ohio Area, edited by Dancey, William S., pp. 7794. Ohio Archaeological Council, Columbus.Google Scholar
Shane, Linda C. K. 1994 Intensity and Rate of Vegetation and Climatic Change in the Ohio Region between 14,000 and 9,000 14C YBP. In The First Discovery of America: Archaeological Evidence of the Early Inhabitants of the Ohio Area, edited by Dancey, William S., pp. 722. Ohio Archaeological Council, Columbus.Google Scholar
Sharp, Madeleine E., Viswanathan, Jayalakshmi, Lanyon, Linda J., and Barton, Jason J. S. 2012 Sensitivity and Bias in Decision-Making under Risk: Evaluating the Perception of Reward, Its Probability and Value. PLoS ONE 7(4):e33460. DOI:10.1371/journal.pone.0033460.CrossRefGoogle ScholarPubMed
Sherman, Christopher E. 1999 Principal Streams and Their Drainage Areas. ODNR Division of Water Resources, Columbus, Ohio.Google Scholar
Shott, Michael J. 2002 Sample Bias in the Distribution and Abundance of Midwestern Fluted Bifaces. Midcontinental Journal of Archaeology 27:89123.Google Scholar
Shott, Michael J. 2004 Representativity of the Midwestern Paleoindian Sample. North American Archaeologist 25:189212.CrossRefGoogle Scholar
Simon, Herbert A. 1956 Rational Choice and the Structure of the Environment. Psychological Review 63:129138.CrossRefGoogle ScholarPubMed
Simons, Donald B. 1997 The Gainey and Butler Sites as Focal Points for Caribou and People. In Caribou and Reindeer Hunters of the Northern Hemisphere, edited by Jackson, Lawrence J. and Thacker, Paul T., pp. 105131. Avebury, Aldershot, United Kingdom.Google Scholar
Steele, James, Adams, Jonathan, and Sluckin, Tim 1998 Modelling Paleoindian Dispersals. World Archaeology 30:286305.CrossRefGoogle Scholar
Stout, Wilbur, and Schoenlaub, R. A. 1945 The Occurrence of Flint in Ohio. Fourth Series—Bulletin 46. State of Ohio Department of Natural Resources, Division of Geological Survey of Ohio, Columbus.Google Scholar
Surface-Evans, Sarah, and White, Devin Alan 2012 An Introduction to the Least Cost Analysis of Social Landscapes. In Least Cost Analysis of Social Landscapes: Archaeological Case Studies, edited by White, Devin Alan and Surface-Evans, Sarah, pp. 17. University of Utah Press, Salt Lake City.Google Scholar
Surovell, Todd A. 2000 Early Paleoindian Women, Children, Mobility, and Fertility. American Antiquity 65:493508.CrossRefGoogle ScholarPubMed
Surovell, Todd A., Finley, Judson Byrd, Smith, Geoffrey M., Jeffrey Brantingham, P., and Kelly, Robert 2009 Correcting Temporal Frequency Distributions for Taphonomic Bias. Journal of Archaeological Science 36:17151724.CrossRefGoogle Scholar
Taleb, Nassim N. 2010 The Black Swan: The Impact of the Highly Improbable. Random House, New York.Google Scholar
Tankersley, Kenneth B. 1994 Was Clovis a Colonizing Population in Eastern North America? In The First Discovery of America: Archaeological Evidence of the Early Inhabitants of the Ohio Area, edited by Dancey, William S., pp. 95116. Ohio Archaeological Council, Columbus.Google Scholar
Tankersley, Kenneth B., Smith, Edward E., and Cochran, Donald R. 1990 Early Paleoindian Land Use, Mobility, and Lithic Exploitation Patterns: An Updated Distribution of Fluted Points in Indiana. North American Archaeologist 11:301319.CrossRefGoogle Scholar
Tobler, Waldo 1993 Three Presentations on Geographical Analysis and Modeling: Non-Isotropic Geographic Modeling Speculations on the Geometry of Geography Global Spatial Analysis. Technical report 93-1. National Center for Geographic Information and Analysis, Santa Barbara, California.Google Scholar
Tolan-Smith, Christopher 2003 The Social Context of Landscape Learning and the Lateglacial-Early Postglacial Recolonization of the British Isles. In Colonization of Unfamiliar Landscapes: The Archaeology of Adaptation, edited by Rockman, Marcy and Steele, James, pp. 116129. Routledge, New York.Google Scholar
Tune, Jesse W. 2016a The Paleoindian and Early Archaic Records in Tennessee: A Review of the Tennessee Fluted Point Survey. Tennessee Archaeologist 8:2441.Google Scholar
Tune, Jesse W. 2016b Characterizing Cumberland Fluted Biface Morphology and Technological Organization. Journal of Archaeological Science: Reports 6:310320.Google Scholar
Tversky, Amos, and Kahneman, Daniel 1992 Advances in Prospect Theory: Cumulative Representation of Uncertainty. Journal of Risk and Uncertainty 5:297323.CrossRefGoogle Scholar
U.S. Geological Survey 2020 National Water Information System. Electronic document, http://waterdata.usgs.gov/nwis/, accessed June 4, 2020.Google Scholar
van der Nat, Dimitry, Tockner, Klement, Edwards, Peter J., Ward, J. V., and Gurnell, Angela M. 2003 Habitat Change in Braided Flood Plains (Tagliamento, NE-Italy). Freshwater Biology 48:17991812.CrossRefGoogle Scholar
van Dommelen, Peter 2014 Moving On: Archaeological Perspectives on Mobility and Migration. World Archaeology 46:477483.CrossRefGoogle Scholar
Vickery, Kent D. 1983 The Flint Sources. In Recent Excavations at the Edwin Harness Mound, Liberty Works, Ross County, Ohio, edited by N'omi, Greber, pp. 7385. Midcontinental Journal of Archaeology Special Publication 5. Kent State University Press, Kent, Ohio.Google Scholar
Webber, Earl E., and Bartlett, William P. 1976 Floods in Ohio—Magnitude and Frequency. USGS Open File Report 76–768. Columbus, Ohio.CrossRefGoogle Scholar
White, Devin Alan 2015 The Basics of Least Cost Analysis for Archaeological Applications. Advances in Archaeological Practice 3:407414.CrossRefGoogle Scholar
Williams, Alan N. 2012 The Use of Summed Radiocarbon Probability Distributions in Archaeology: A Review of Methods. Journal of Archaeological Science 39:578589.CrossRefGoogle Scholar
Winterhalder, Bruce 1986 Diet Choice, Risk, and Food Sharing in a Stochastic Environment. Journal of Anthropological Archaeology 5:369392.CrossRefGoogle Scholar
Winterhalder, Bruce 2007 Risk and Decision-Making. In Oxford Handbook of Evolutionary Psychology, edited by Dunbar, R. I. M. and Barrett, Louise, pp. 443446. Oxford University Press, Oxford.Google Scholar
Winterhalder, Bruce, Lu, Flora, and Tucker, Bram 1999 Risk-Senstive Adaptive Tactics: Models and Evidence from Subsistence Studies in Biology and Anthropology. Journal of Archaeological Research 7:301348.CrossRefGoogle Scholar
Yu, Zicheng 2000 Ecosystem Response to Late Glacial and Early Holocene Climate Oscillations in the Great Lakes Region of North America. Quaternary Science Reviews 19:17231747.CrossRefGoogle Scholar
Yu, Zicheng, and Eicher, Ulrich 2001 Three Amphi-Atlantic Century-Scale Cold Events during the Bølling-Allerød Warm Period. Géographie Physique et Quaternaire 55:171179.CrossRefGoogle Scholar