Skip to main content
Log in

Design Space and Cultural Transmission: Case Studies from Paleoindian Eastern North America

  • Published:
Journal of Archaeological Method and Theory Aims and scope Submit manuscript

Abstract

Tool design is a cultural trait—a term long used in anthropology as a unit of transmittable information that encodes particular behavioral characteristics of individuals or groups. After they are transmitted, cultural traits serve as units of replication in that they can be modified as part of a cultural repertoire through processes such as recombination, loss, or partial alteration. Artifacts and other components of the archaeological record serve as proxies for studying the transmission (and modification) of cultural traits, provided there is analytical clarity in defining and measuring whatever it is that is being transmitted. Our interest here is in tool design, and we illustrate how to create analytical units that allow us to map tool-design space and to begin to understand how that space was used at different points in time. We first introduce the concept of fitness landscape and impose a model of cultural learning over it, then turn to four methods that are useful for the analysis of design space: paradigmatic classification, phylogenetic analysis, distance graphs, and geometric morphometrics. Each method builds on the others in logical fashion, which allows creation of testable hypotheses concerning cultural transmission and the evolutionary processes that shape it, including invention (mutation), selection, and drift. For examples, we turn to several case studies that focus on Early Paleoindian–period projectile points from eastern North America, the earliest widespread and currently recognizable remains of hunter–gatherers in late Pleistocene North America.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16

Similar content being viewed by others

Notes

  1. Our landscape here is strictly impressionistic and intended only as an example. Realistic three-dimensional maps of decision making that incorporate fitness peaks are extremely complicated to model (Brock et al. 2014).

  2. Although the three characters are constructed from a continuous variable—length—we can divide it into any number of bins.

  3. Technically, cladistics is a particular method for creating hypotheses of evolutionary relationships (Brinkman and Leipe 2001), but we follow what has become common practice and refer to several phylogenetic methods collectively as cladistics. These include maximum likelihood and Bayesian Markov Chain Monte Carlo, both of which calculate probabilities through reference to an explicit evolutionary model from which the data are assumed to be distributed identically (Kolaczkowski and Thornton 2004).

  4. O’Brien et al. (2014) note that the database consists of 1813 specimens. That number is a typographical error.

  5. Because it is much easier to use class abbreviations than it is to write out class definitions (the number strings)—not to mention easier to remember—we use the abbreviations. Note that the abbreviations are based on commonly used type names. In each case, the type names were taken directly from the literature in which the specimens were illustrated. For example, Class DAQS contains six specimens, at least one of which was originally referred to as a Dalton (D) point, at least one as an Arkabutla (A) point, at least one as a Quad (Q) point, and at least one as a Simpson (S) point. Echoing our discussion in the text, this ought to give us pause the next time we think about using traditional projectile-point types as analytical units.

  6. The number of specimens changes from character to character because we were conservative in which specimens to include. All 1113 points included in the overall sample were complete in terms of assigning particular character states to them, but they might have had small pieces missing that precluded exact measurements and thus were excluded from the analysis shown in Fig. 11.

  7. One of our recent studies (Eren et al. 2015b) focused on Paleoindian-point richness in eastern North America, with richness being the number of classes in our database per region. When the Southeast as a whole is compared statistically to the Northeast, Mason’s (1962) prediction is upheld. Further, when we divide the Southeast into the Upper Southeast and the Lower Southeast, the former region exhibits greater class richness than either of the other two regions. The Upper Southeast contains portions of major waterways such as the Missouri, Mississippi, and Ohio rivers.

  8. Although not presented here, analysis of data generated by binning specimens by longitude showed no patterning. This parallels the findings of Smith et al. (2015), who note that in their study latitude was associated more strongly with shape than longitude was.

References

  • Anderson, D. G. (1990). The Paleoindian colonization of eastern North America: a view from the southeastern United States. In K. B. Tankersely & B. L. Isaac (Eds.), Early Paleoindian economies of eastern North America (pp. 163–216). Greenwich, CT: JAI Press.

    Google Scholar 

  • Anderson, D. G. (1995). Paleoindian interaction networks in the Eastern Woodlands. In M. S. Nassaney & K. E. Sassaman (Eds.), Native American interaction: multiscalar analyses and interpretations in the Eastern Woodlands (pp. 1–26). Knoxville: University of Tennessee Press.

    Google Scholar 

  • Anderson, D. G., & Faught, M. K. (1998). The distribution of fluted Paleoindian projectile points: update 1998. Archaeology of Eastern North America, 26, 163–187.

    Google Scholar 

  • Anderson, D. G., & Faught, M. K. (2000). Palaeoindian artifact distributions: evidence and implications. Antiquity, 74, 507–513.

    Google Scholar 

  • Anderson, D. G., O’Steen, L. D., & Sassaman, K. E. (1996). Chronological considerations. In D. G. Anderson & K. E. Sassaman (Eds.), The Paleoindian and Early Archaic Southeast (pp. 3–15). Tuscaloosa: University of Alabama Press.

    Google Scholar 

  • Anderson, D. G., Miller, D. S., Yerka, S. J., Gillam, J. C., Johanson, D. T., Goodyear, A. C., & Smallwood, A. M. (2010). Paleoindian database of the Americas 2010: current status and findings. Archaeology of Eastern North America, 38, 63–90.

    Google Scholar 

  • Anderson, D. G., Smallwood, A. M., & Miller, D. S. (2015). Pleistocene human settlement in the southeastern United States: current evidence and future directions. PaleoAmerica, 1, 7–51.

    Article  Google Scholar 

  • Andrews, B. N., Knell, E. J., & Eren, M. I. (2015). The three lives of a uniface. Journal of Archaeological Science, 54, 228–236.

    Article  Google Scholar 

  • Aoki, K., & Feldman, M. W. (2014). Evolution of learning strategies in temporally and spatially variable environments: a review of theory. Theoretical Population Biology, 91, 3–19.

    Article  Google Scholar 

  • Aoki, K., & Mesoudi, A. (2015). Introduction to “Learning strategies and cultural evolution during the Palaeolithic.” In A. Mesoudi & K. Aoki (Eds.), Learning strategies and cultural evolution during the Palaeolithic (pp. 1–8). Tokyo: Springer.

    Google Scholar 

  • Archer, W., & Braun, D. R. (2010). Variability in bifacial technology at Elandsfontein, Western Cape, South Africa: a geometric morphometric approach. Journal of Archaeological Science, 37, 201–209.

    Article  Google Scholar 

  • Archibald, J. K., Mort, M. E., & Crawford, D. J. (2003). Bayesian inference of phylogeny: a non-technical primer. Taxon, 52, 187–191.

    Article  Google Scholar 

  • Arnold, M. L. (1997). Natural hybridization and evolution. New York: Oxford University Press.

    Google Scholar 

  • Atkisson, C., O’Brien, M. J., & Mesoudi, A. (2012). Adult learners in a novel environment use prestige-biased social learning. Evolutionary Psychology, 10, 519–537.

    Article  Google Scholar 

  • Aunger, R. (2002). The electric meme: a new theory of how we think. New York: Free Press.

    Google Scholar 

  • Baxter, M. J., & Cool, H. E. M. (2010). Detecting modes in low-dimensional archaeological data. Journal of Archaeological Science, 37, 2379–2385.

    Article  Google Scholar 

  • Beck, C., & Jones, G. T. (1989). Bias and archaeological classification. American Antiquity, 54, 244–263.

    Article  Google Scholar 

  • Beck, C., & Jones, G. T. (2007). Early Paleoarchaic point morphology and chronology. In K. Graf & D. N. Schmitt (Eds.), Paleoindian or Paleoarchaic? Great Basin human ecology at the Pleistocene–Holocene transition (pp. 23–41). Salt Lake City: University of Utah Press.

    Google Scholar 

  • Bement, L. C., & Carter, B. J. (2015). From mammoth to bison: changing Clovis prey availability at the end of the Pleistocene. In A. M. Smallwood & T. A. Jennings (Eds.), Clovis: on the edge of a new understanding (pp. 263–275). College Station: Texas A&M University Press.

    Google Scholar 

  • Bentley, R. A., & O’Brien, M. J. (2011). The selectivity of cultural learning and the tempo of cultural evolution. Journal of Evolutionary Psychology, 9, 125–141.

    Article  Google Scholar 

  • Bentley, R. A., Lipo, C. P., Herzog, H. A., & Hahn, M. W. (2007). Regular rates of popular culture change reflect random copying. Evolution and Human Behavior, 28, 151–158.

    Article  Google Scholar 

  • Bentley, R. A., Earls, M., & O’Brien, M. J. (2011a). I’ll have what she’s having: mapping social behavior. Cambridge, MA: MIT Press.

    Google Scholar 

  • Bentley, R. A., O’Brien, M. J., & Ormerod, P. (2011b). Quality versus mere popularity: a conceptual map for understanding human behavior. Mind and Society, 10, 181–191.

    Article  Google Scholar 

  • Bentley, R. A., O’Brien, M. J., & Brock, W. A. (2014). Mapping collective behavior in the big-data era. Behavioral and Brain Sciences, 37, 63–119.

    Article  Google Scholar 

  • Bettinger, R. L., & Eerkens, J. W. (1997). Evolutionary implications of metrical variation in Great Basin projectile points. In C. M. Barton & G. A. Clark (Eds.), Rediscovering Darwin: evolutionary theory and archeological explanation (pp. 177–191). Washington, DC: American Anthropological Association, Archeological Papers 7.

    Google Scholar 

  • Bettinger, R. L., & Eerkens, J. (1999). Point typologies, cultural transmission, and the spread of bow-and-arrow technology in the prehistoric Great Basin. American Antiquity, 64, 231–242.

    Article  Google Scholar 

  • Bingham, P. M., Souza, J., & Blitz, J. H. (2013). Social complexity and the bow in the prehistoric North American record. Evolutionary Anthropology, 22, 81–88.

    Article  Google Scholar 

  • Bleed, P. (1986). The optimal design of hunting weapons: maintainability or reliability. American Antiquity, 45, 4–20.

    Google Scholar 

  • Bookstein, L. (1991). Morphometric tools for landmark data: geometry and biology. Cambridge: Cambridge University Press.

    Google Scholar 

  • Borgerhoff Mulder, M., Nunn, C. L., & Towner, M. C. (2006). Cultural macroevolution and the transmission of traits. Evolutionary Anthropology, 15, 52–64.

    Article  Google Scholar 

  • Boulanger, M. T., & Hudson, C. M. (2012). Assessment of the gripability of textured ceramic surfaces. American Antiquity, 77, 293–302.

    Article  Google Scholar 

  • Boulanger, M. T., Buchanan, B., O’Brien, M. J., Redmond, B. G., Glascock, M. D., & Eren, M. 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, U.S.A. Journal of Archaeological Science, 53, 550–558.

    Article  Google Scholar 

  • Bowern, C. (2012). The riddle of Tasmanian languages. Proceedings of the Royal Society B, 279, 4590–4595.

    Article  Google Scholar 

  • Boyd, R., & Richerson, P. J. (1985). Culture and the evolutionary process. London: University of Chicago Press.

    Google Scholar 

  • Boyd, R., & Richerson, P. J. (1995). Why does culture increase human adaptability? Ethology and Sociobiology, 16, 125–143.

    Article  Google Scholar 

  • Boyd, R., & Richerson, P. J. (2010). Transmission coupling mechanisms: cultural group selection. Philosophical Transactions of the Royal Society B, 365, 3787–3790.

    Article  Google Scholar 

  • Boyd, R., Richerson, P. J., Borgerhoff-Mulder, M., & Durham, W. H. (1997). Are cultural phylogenies possible? In P. Weingart, P. J. Richerson, S. D. Mitchell, & S. Maasen (Eds.), Human by nature: between biology and the social sciences (pp. 355–386). Mahwah, NJ: Erlbaum.

    Google Scholar 

  • Bradley, J. W., Spiess, A., Boisvert, R. A., & Boudreau, J. (2008). What’s the point? Modal forms and attributes of Paleoindian bifaces in the New England–Maritimes Region. Archaeology of Eastern North America, 36, 119–172.

    Google Scholar 

  • Bradley, B. A., Collins, M. B., & Hemmings, A. (2010). Clovis technology. Ann Arbor, MI: International Monographs in Prehistory.

    Google Scholar 

  • Brantingham, P. J., & Kuhn, S. L. (2001). Constraints on Levallois core technology: a mathematical model. Journal of Archaeological Science, 28, 747–761.

    Article  Google Scholar 

  • Brew, J. O. (1946). The archaeology of Alkali Ridge, southeastern Utah. Peabody Museum of American Archaeology and Ethnology, Papers 21. Cambridge, MA.

  • Brinkman, F. S. L., & Leipe, D. D. (2001). Phylogenetic analysis. In A. D. Baxevanis & B. F. F. Ouellette (Eds.), Bioinformatics: a practical guide to the analysis of genes and proteins (2nd ed., pp. 323–358). New York: Wiley.

    Chapter  Google Scholar 

  • Brock, W. A., Bentley, R. A., O’Brien, M. J., & Caido, C. C. S. (2014). Estimating a path through a map of decision making. PLoS ONE, 9(11), e111022.

    Article  Google Scholar 

  • Brower, A. V. Z. (2009). Science as a pattern. Systematics and Biodiversity, 7, 345–346.

    Article  Google Scholar 

  • Brower, A. V. Z., & de Pinna, M. C. C. (2012). Homology and errors. Cladistics, 28, 529–538.

    Article  Google Scholar 

  • Brown, J. H., & Lomolino, M. V. (1998). Biogeography (2nd ed.). Sunderland, MA: Sinauer.

    Google Scholar 

  • Bryant, D., & Moulton, V. (2002). Neighbor-Net, an agglomerative algorithm for the construction of phylogenetic networks. In R. Guigo & D. Gusfield (Eds.), Algorithms in bioinformatics: second international workshop (pp. 375–391). Berlin: Springer.

    Chapter  Google Scholar 

  • Bryant, D., Filimon, F., & Gray, R. D. (2005). Untangling our past: languages, trees, splits and networks. In R. Mace, C. J. Holden, & S. Shennan (Eds.), The evolution of cultural diversity: a phylogenetic approach (pp. 69–85). London: University College London Press.

    Google Scholar 

  • Buchanan, B., & Collard, M. (2010). A geometric morphometrics-based assessment of blade shape differences among Paleoindian projectile point types from western North America. Journal of Archaeological Science, 37, 350–359.

    Article  Google Scholar 

  • Buchanan, B., & Hamilton, M. J. (2009). A formal test of the origin of variation in North American Early Paleoindian projectile points. American Antiquity, 74, 279–298.

    Google Scholar 

  • Buchanan, B., Collard, M., Hamilton, M. J., & O’Brien, M. J. (2011). Points and prey: an evaluation of the hypothesis that prey size predicts early Paleoindian projectile point form. Journal of Archaeological Science, 38, 852–864.

    Article  Google Scholar 

  • Buchanan, B., Kilby, J. D., Huckell, B. B., O’Brien, M. J., & Collard, M. (2012). A morphometric assessment of the intended function of cached Clovis points. PLoS ONE, 7(2), e30530.

    Article  Google Scholar 

  • Buchanan, B., O’Brien, M. J., & Collard, M. (2014). Continent-wide or region-specific? A geometric-morphometrics-based assessment of variation in Clovis point shape. Archaeological and Anthropological Sciences, 6, 145–162.

    Article  Google Scholar 

  • Buchanan, B., Eren, M. I., Boulanger, M. T., & O’Brien, M. J. (2015). Size, shape, scars, and spatial patterning: a quantitative assessment of late Pleistocene (Clovis) point resharpening. Journal of Archaeological Science: Reports, 3, 11−21.

  • Buikstra, J. E., Konigsberg, L., & Bullington, J. (1986). Fertility and the development of agriculture in the prehistoric Midwest. American Antiquity, 51, 528–546.

    Article  Google Scholar 

  • Cardillo, M. (2010). Some applications of geometric morphometrics to archaeology. In A. M. T. Elewa (Ed.), Morphometrics for nonmorphometricians (pp. 325–341). Berlin: Springer-Verlag.

    Chapter  Google Scholar 

  • Catalano, S. A., & Goloboff, P. A. (2012). Simultaneously mapping and superimposing landmark configurations with parsimony as optimality criterion. Systematic Biology, 61, 392–400.

    Article  Google Scholar 

  • Catalano, S. A., Goloboff, P. A., & Giannini, N. P. (2010). Phylogenetic morphometrics (I): the use of landmark data in a phylogenetic framework. Cladistics, 26, 539–549.

    Article  Google Scholar 

  • Charlin, J., & González-José, R. (2012). Size and shape variation in late Holocene projectile points of southern Patagonia: a geometric morphometric study. American Antiquity, 77, 221–242.

    Article  Google Scholar 

  • Claidière, N., & André, J.-B. (2012). The transmission of genes and culture: a questionable analogy. Evolutionary Biology, 39, 12–24.

    Article  Google Scholar 

  • Cochrane, E. E. (2013). Quantitative phylogenetic analyses of Lapita decoration in Near and Remote Oceania. In G. Summerhayes & H. Buckley (Eds.), Pacific archaeology: documenting the past 50,000 years (pp. 17–42). Dunedin, New Zealand: University of Otago.

    Google Scholar 

  • Cochrane, E. E., & Lipo, C. P. (2010). Phylogenetic analyses of Lapita decoration do not support branching evolution or regional population structure during colonization of Remote Oceania. Philosophical Transactions of the Royal Society B, 365, 3889–3902.

    Article  Google Scholar 

  • Collard, M., Shennan, S. J., & Tehrani, J. J. (2006). Branching, blending and the evolution of cultural similarities and differences among human populations. Evolution and Human Behavior, 27, 169–184.

    Article  Google Scholar 

  • Costa, A. G. (2010). A geometric morphometric assessment of plan shape in bone and stone Acheulean bifaces from the middle Pleistocene site of Castel di Guido, Latium, Italy. In S. J. Lycett & P. R. Chauhan (Eds.), New perspectives on old stones: analytical approaches to Paleolithic technologies (pp. 23–59). New York: Springer.

    Chapter  Google Scholar 

  • Cotter, J. L. (1937). The occurrence of flints and extinct animals in pluvial deposits near Clovis, New Mexico: part IV, report on excavation at the gravel pit, 1936. Proceedings of the Academy of Natural Sciences of Philadelphia, 89, 1–16.

    Google Scholar 

  • Cotter, J. L. (1938). The occurrence of flints and extinct animals in pluvial deposits near Clovis, New Mexico: part VI, report on field season of 1937. Proceedings of the Academy of Natural Sciences of Philadelphia, 90, 113–117.

    Google Scholar 

  • Coward, F., Shennan, S., Colledge, S., Conolly, J., & Collard, M. (2008). The spread of Neolithic plant economies from the Near East to northwest Europe: a phylogenetic analysis. Journal of Archaeological Science, 35, 42–56.

    Article  Google Scholar 

  • Cox, S. L. (1986). A re-analysis of the Shoop site. Archaeology of Eastern North America, 14, 101–170.

    Google Scholar 

  • Crabtree, D. E. (1966). A stoneworker’s approach to analyzing and replicating the Lindenmeier Folsom. Tebiwa, 9, 3–39.

    Google Scholar 

  • Crema, E. R., Kerig, T., & Shennan, S. (2014). Culture, space, and metapopulation: a simulation-based study for evaluating signals of blending and branching. Journal of Archaeological Science, 43, 289–298.

    Article  Google Scholar 

  • Currie, T. E., & Mace, R. (2011). Mode and tempo in the evolution of socio-political organization: reconciling ‘Darwinian’ and ‘Spencerian’ approaches in anthropology. Philosophical Transactions of the Royal Society B, 366, 1108–1117.

    Article  Google Scholar 

  • Currie, T. E., Greenhill, S. J., Gray, R. D., Hasegawa, T., & Mace, R. (2010). Rise and fall of political complexity in island South-East Asia and the Pacific. Nature, 467, 801–804.

    Article  Google Scholar 

  • Dagan, T., & Martin, W. (2007). Ancestral genome sizes specify the minimum rate of lateral gene transfer during prokaryote evolution. Proceedings of the National Academy of Sciences, 104, 870–875.

    Article  Google Scholar 

  • Darwent, J., & O’Brien, M. J. (2006). Using cladistics to construct lineages of projectile points from northeastern Missouri. In C. P. Lipo, M. J. O’Brien, M. Collard, & S. J. Shennan (Eds.), Mapping our ancestors: phylogenetic approaches in anthropology and prehistory (pp. 185–208). New York: Aldine.

    Google Scholar 

  • Denton, G. H., Anderson, R. F., Toggweiler, J. R., Edwards, R. L., Schaeer, J. M., & Putnam, A. E. (2010). The last glacial termination. Science, 328, 1652–1656.

    Article  Google Scholar 

  • Doolittle, W. F. (1999). Phylogenetic classification and the universal tree. Science, 284, 2124–2128.

    Article  Google Scholar 

  • Driver, H. E. (1973). Cultural diffusion. In R. Naroll & F. Naroll (Eds.), Main currents in cultural anthropology (pp. 157–183). New York: Appleton-Century-Crofts.

    Google Scholar 

  • Dunnell, R. C. (1971). Systematics in prehistory. New York: Free Press.

    Google Scholar 

  • Dunnell, R. C. (1986). Methodological issues in Americanist artifact classification. In M. B. Schiffer (Ed.), Advances in archaeological method and theory (vol. 9) (pp. 149–207). Orlando, FL: Academic Press.

    Google Scholar 

  • Dunnell, R. C. (1989). Aspects of the application of evolutionary theory in archaeology. In C. C. Lamberg-Karlovsky (Ed.), Archaeological thought in America (pp. 35–49). Cambridge: Cambridge University Press.

    Chapter  Google Scholar 

  • Eerkens, J. W., & Lipo, C. P. (2005). Cultural transmission, copying errors, and the generation of variation in material culture and the archaeological record. Journal of Anthropological Archaeology, 24, 316–334.

    Article  Google Scholar 

  • Eerkens, J. W., Bettinger, R. L., & McElreath, R. (2006). Cultural transmission, phylogenetics, and the archaeological record. In C. P. Lipo, M. J. O’Brien, M. Collard, & S. J. Shennan (Eds.), Mapping our ancestors: phylogenetic approaches in anthropology and prehistory (pp. 169–183). New York: Aldine.

    Google Scholar 

  • Eerkens, J. W., Bettinger, R. L., & Richerson, P. J. (2014). Cultural transmission theory and hunter–gatherer archaeology. In V. Cummings, P. Jordan, & M. Zvelebil (Eds.), The Oxford handbook of the archaeology and anthropology of hunter–gatherers (pp. 1127–1142). Oxford: Oxford University Press.

    Google Scholar 

  • Efron, B., Halloran, E., & Holmes, S. (1996). Bootstrap confidence levels for phylogenetic trees. Proceedings of the National Academy of Sciences, 93, 7085–7090.

    Article  Google Scholar 

  • Ellis, C. (2004). Understanding “Clovis” fluted point variability in the Northeast: a perspective from the Debert site, Nova Scotia. Canadian Journal of Archaeology, 28, 205–253.

    Google Scholar 

  • Ellis, C., & Deller, D. B. (1997). Variability in the archaeological record of northeastern Early Paleoindians: a view from southern Ontario. Archaeology of Eastern North America, 25, 1–30.

    Google Scholar 

  • Endler, J. A. (1998). The place of hybridization in evolution. Evolution, 52, 640–644.

    Article  Google Scholar 

  • Enquist, M., Eriksson, K., & Ghirlanda, S. (2007). Critical social learning: a solution to Rogers’ paradox of non-adaptive culture. American Anthropologist, 109, 727–734.

    Article  Google Scholar 

  • Eren, M. I. (2012). Were unifacial tools regularly hafted by Clovis foragers in the North American Lower Great Lakes region? An empirical test of edge class richness and attribute frequency among distal, proximal, and lateral tool-sections. Journal of Ohio Archaeology, 2, 1–15.

    Google Scholar 

  • Eren, M. I., & Desjardines, A. (2015). Flaked stone tools of Pleistocene colonizers: overshot flaking at the Redwing site, Ontario. In A. M. Smallwood & T. A. Jennings (Eds.), Clovis: on the edge of a new understanding (pp. 109–120). College Station: Texas A&M University Press.

    Google Scholar 

  • Eren, M. I., & Lycett, S. J. (2012). Why Levallois? A morphometric comparison of experimental ‘preferential’ Levallois flakes versus debitage flakes. PLoS ONE, 7(1), e29273.

    Article  Google Scholar 

  • Eren, M. I., Bradley, B. A., & Sampson, C. G. (2011a). Middle Paleolithic skill level and the individual knapper: an experiment. American Antiquity, 76, 229–251.

    Article  Google Scholar 

  • Eren, M. I., Lycett, S. J., Roos, C. I., & Sampson, C. G. (2011b). Toolstone constraints on knapping skill: Levallois reduction with two different rock types. Journal of Archaeological Science, 38, 2731–2739.

    Article  Google Scholar 

  • Eren, M. I., Patten, R. J., O’Brien, M. J., & Meltzer, D. J. (2013). Refuting the technological cornerstone of the Ice-Age Atlantic crossing hypothesis. Journal of Archaeological Science, 40, 2934–2941.

    Article  Google Scholar 

  • Eren, M. I., Roos, C. I., Story, B. A., von Cramon-Taubadel, N., & Lycett, S. J. (2014). The role of raw material differences in stone tool shape variation: an experimental assessment. Journal of Archaeological Science, 49, 472–487.

    Article  Google Scholar 

  • Eren, M. I., Buchanan, B., & O’Brien, M. J. (2015a). Social learning and technological evolution during Clovis colonization of the New World. Journal of Human Evolution, 80, 159–170.

    Article  Google Scholar 

  • Eren, M. I., Chao, A., Chiu, C., Colwell, R. K., Buchanan, B., Boulanger, M. T., Darwent, J., O’Brien, M. J. (2015b) Statistical analysis of paradigmatic class richness supports greater Paleoindian projectile-point diversity in the Southeast. American Antiquity (in press).

  • Farris, J. S. (1989a). The retention index and the rescaled consistency index. Cladistics, 5, 417–419.

    Article  Google Scholar 

  • Farris, J. S. (1989b). The retention index and homoplasy excess. Systematic Zoology, 18, 1–32.

    Google Scholar 

  • Farris, J. S. (2014). Homology and misdirection. Cladistics, 30, 555–561.

    Article  Google Scholar 

  • Faught, M. K. (2006). Paleoindian archaeology in Florida and Panama: two Circumgulf regions exhibiting waisted lanceolate projectile points. In J. Morrow & C. Gnecco (Eds.), Ice Age occupations of the Americas: a hemispheric perspective (pp. 164–183). Gainesville: University Press of Florida.

    Google Scholar 

  • Felsenstein, J. (1985). Confidence limits on phylogenies: an approach using the bootstrap. Evolution, 39, 783–791.

    Article  Google Scholar 

  • Fiedel, S. J. (2015). The Clovis-era radiocarbon plateau. In A. M. Smallwood & T. A. Jennings (Eds.), Clovis: on the edge of a new understanding (pp. 11–19). College Station: Texas A&M University Press.

    Google Scholar 

  • Figgins, J. D. (1933). A further contribution to the antiquity of man in America. Colorado Museum of Natural History Proceedings No. 12. Denver.

  • García Rivero, D., & O’Brien, M. J. (2014). Phylogenetic analysis shows that Neolithic slate plaques from the southwestern Iberian Peninsula are not genealogical recording systems. PLoS ONE, 9(2), e88296.

    Article  Google Scholar 

  • Gingerich, J. A. M., Sholts, S. B., Wärmländer, S. K. T. S., & Stanford, D. (2014). Fluted point manufacture in eastern North America: an assessment of form and technology using traditional metrics and 3D digital morphometrics. World Archaeology, 46, 101–122.

    Article  Google Scholar 

  • Goebel, T., Waters, M. R., & O’Rourke, D. H. (2008). The late Pleistocene dispersal of modern humans in the Americas. Science, 319, 1497–1502.

    Article  Google Scholar 

  • Goloboff, P. A. (2003). Parsimony, likelihood, and simplicity. Cladistics, 19, 91–103.

    Article  Google Scholar 

  • Goloboff, P. A., & Catalano, S. A. (2011). Phylogenetic morphometrics (II): algorithms for landmark optimization. Cladistics, 27, 42–51.

    Article  Google Scholar 

  • Gould, S. J. (1991). The disparity of the Burgess Shale arthropod fauna and the limits of cladistic analysis: why we must strive to quantify morphospace. Paleobiology, 17, 411–423.

    Google Scholar 

  • Gould, S. J. (2002). The structure of evolutionary theory. Belknap: Cambridge, MA.

    Google Scholar 

  • Graf, K. E., Ketron, C. V., & Waters, M. R. (Eds.). (2014). Paleoamerican odyssey. College Station: Texas A&M University Press.

    Google Scholar 

  • Gray, R. D., & Atkinson, Q. D. (2003). Language-tree divergence times support the Anatolian theory of Indo-European origin. Nature, 426, 435–439.

    Article  Google Scholar 

  • Gray, R. D., Drummond, A. J., & Greenhill, S. J. (2009). Language phylogenies reveal expansion pulses and pauses in Pacific settlement. Science, 323, 479–483.

    Article  Google Scholar 

  • Gray, R. D., Bryant, D., & Greenhill, S. J. (2010). On the shape and fabric of human history. Philosophical Transactions of the Royal Society B, 365, 3923–3933.

    Article  Google Scholar 

  • Hamilton, M. J. (2008). Quantifying Clovis dynamics: conforming theory with models and data across scales. PhD dissertation, Department of Anthropology, University of New Mexico. Albuquerque.

  • Hamilton, M. J., & Buchanan, B. (2009). The accumulation of stochastic copying errors causes drift in culturally transmitted technologies: quantifying Clovis evolutionary dynamics. Journal of Anthropological Archaeology, 28, 55–69.

    Article  Google Scholar 

  • Hamming, R. (1980). Coding and information theory. Upper Saddle River, NJ: Prentice-Hall.

    Google Scholar 

  • Hart, J. P., & Brumbach, H. J. (2003). The death of Owasco. American Antiquity, 68, 737–752.

    Article  Google Scholar 

  • Haynes, C. V. (1964). Fluted projectile points: their age and dispersion. Science, 145, 1408–1413.

    Article  Google Scholar 

  • Haynes, C. V. (1983). Fluted points in the East and West. Archaeology of Eastern North America, 11, 24–26.

    Google Scholar 

  • Heggarty, P., Maguire, W., & McMahon, A. (2010). Splits or waves? Trees or webs? How divergence measures and network analysis can unravel language histories. Philosophical Transactions of the Royal Society B, 365, 3829–3843.

    Article  Google Scholar 

  • Hennig, W. (1966). Phylogenetic systematics. Urbana: University of Illinois Press.

    Google Scholar 

  • Henrich, J. (2001). Cultural transmission and the diffusion of innovations: adoption dynamics indicate that biased cultural transmission is the predominate force in behavioral change. American Anthropologist, 103, 992–1013.

    Article  Google Scholar 

  • Henrich, J. (2004). Demography and cultural evolution: why adaptive cultural processes produced maladaptive losses in Tasmania. American Antiquity, 69, 197–214.

    Article  Google Scholar 

  • Henrich, J. (2006). Understanding cultural evolutionary models: a reply to Read’s critique. American Antiquity, 71, 771–782.

    Article  Google Scholar 

  • Henrich, J., & Boyd, R. (1998). The evolution of conformist transmission and the emergence of between-group differences. Evolution and Human Behavior, 19, 215–241.

    Article  Google Scholar 

  • Henrich, J., & Broesch, J. (2011). On the nature of cultural transmission networks: evidence from Fijian villages for adaptive learning biases. Philosophical Transactions of the Royal Society, 366, 1139–1148.

    Article  Google Scholar 

  • Henrich, J., & Gil-White, F. J. (2001). The evolution of prestige: freely conferred deference as a mechanism for enhancing the benefits of cultural transmission. Evolution and Human Behavior, 22, 165–196.

    Article  Google Scholar 

  • Hester, J. J. (1972). Blackwater locality no. 1: a stratified Early Man site in eastern New Mexico. Fort Burgwin Research Center Publication, no. 8. Ranchos de Taos, NM.

  • Heyes, C. M. (1994). Social learning in animals: categories and mechanisms. Biological Reviews, 69, 207–231.

    Article  Google Scholar 

  • Hoek, W. Z. (2008). The last glacial–interglacial transition. Episodes, 31, 226–229.

    Google Scholar 

  • Holden, C. J., & Mace, R. (2003). Spread of cattle pastoralism led to the loss of matriliny in Africa: a co-evolutionary analysis. Proceedings of the Royal Society B, 270, 2425–2433.

    Article  Google Scholar 

  • Hoppitt, W., & Laland, K. N. (2013). Social learning: an introduction to mechanisms, methods, and models. Princeton, NJ: Princeton University Press.

    Book  Google Scholar 

  • Howard, C. D. (1990). The Clovis point: characteristics and type description. Plains Anthropologist, 35, 255–262.

    Google Scholar 

  • Hull, D. L. (1981). Units of evolution: a metaphysical essay. In U. J. Jenson & R. Harré (Eds.), The philosophy of evolution (pp. 23–44). New York: St. Martin’s Press.

    Google Scholar 

  • Huson, D. H., & Bryant, D. (2006). Application of phylogenetic networks in evolutionary studies. Molecular Biology and Evolution, 23, 254–267.

    Article  Google Scholar 

  • Iovita, R., & McPherron, S. P. (2011). The handaxe reloaded: a morphometric reassessment of Acheulian and Middle Paleolithic handaxes. Journal of Human Evolution, 61, 61–74.

    Article  Google Scholar 

  • Jennings, T. A., & Waters, M. R. (2014). Pre-Clovis lithic technology at the Debra L. Friedkin site, Texas: comparisons to Clovis through site-level behavior, technological trait-list and cladistics analyses. American Antiquity, 79, 25–44.

    Article  Google Scholar 

  • Jordan, P. (2015). Technology as human social tradition: cultural transmission among hunter–gatherers. Berkeley: University of California Press.

    Google Scholar 

  • Jordan, P., & Shennan, S. J. (2003). Cultural transmission, language and basketry traditions amongst the California Indians. Journal of Anthropological Archaeology, 22, 42–74.

    Article  Google Scholar 

  • Jordan, F. M., Gray, R. D., Greenhill, S. J., & Mace, R. (2009). Matrilocal residence is ancestral in Austronesian societies. Proceedings of the Royal Society B, 276, 1957–1964.

    Article  Google Scholar 

  • Justice, N. D. (1987). Stone Age spear and arrow points of the midcontinental and eastern United States: a modern survey and reference. Bloomington: Indiana University Press.

    Google Scholar 

  • Kameda, T., & Nakanishi, D. (2002). Cost-benefit analysis of social/cultural learning in a nonstationary uncertain environment: an evolutionary simulation and an experiment with human subjects. Evolution and Human Behavior, 23, 373–393.

    Article  Google Scholar 

  • Kane, D. (1996). Local hillclimbing on an economic landscape. Santa Fe Institute Working Paper 96-08-065. Santa Fe, NM.

  • Kauffman, S. (1995). At home in the universe: the search for laws of self-organization and complexity. Oxford: Oxford University Press.

    Google Scholar 

  • Kauffman, S., Lobo, J., & Macready, W. (2000). Optimal search on a technology landscape. Journal of Economic Behavior and Organization, 43, 141–166.

    Article  Google Scholar 

  • Kelly, R. L., & Todd, L. C. (1988). Coming into the country: Early Paleoindian hunting and mobility. American Antiquity, 53, 231–244.

    Article  Google Scholar 

  • Kempe, M., Lycett, S. J., & Mesoudi, A. (2012). An experimental test of the accumulated copying error model of cultural mutation for Acheulean handaxe size. PLoS ONE, 7(11), e48333.

    Article  Google Scholar 

  • Kempe, M., Lycett, S. J., & Mesoudi, A. (2014). Cultural differences and cumulative culture: parameterizing the differences between human and nonhuman culture. Journal of Theoretical Biology, 359, 29–36.

    Article  Google Scholar 

  • Key, A. J. (2013). Applied force as a determining factor in lithic use-wear accrual: an experimental investigation of its validity as a method with which to infer hominin upper limb biomechanics. Lithic Technology, 38, 32–45.

    Article  Google Scholar 

  • Key, A. J., & Lycett, S. J. (2015). Edge angle as a variably influential factor in flake cutting efficiency: an experimental investigation of its relationship with tool size and loading. Archaeometry. doi: 10.1111/arcm.12140.

  • Kluge, A. G., & Farris, J. S. (1969). Quantitative phyletics and the evolution of anurans. Systematic Zoology, 18, 1–32.

    Article  Google Scholar 

  • Knappett, C. (Ed.). (2013). Network analysis in archaeology: new approaches to regional integration. London: Oxford University Press.

    Google Scholar 

  • Kolaczkowski, B., & Thornton, J. W. (2004). Performance of maximum parsimony and likelihood phylogenetics when evolution is heterogeneous. Nature, 431, 980–984.

    Article  Google Scholar 

  • Kuhn, S. L. (2012). Emergent patterns of creativity and innovation in early technologies. In S. Elias (Ed.), Origins of human innovation and creativity (pp. 69–87). Amsterdam: Elsevier.

    Chapter  Google Scholar 

  • Kvitek, D. J., & Sherlock, G. (2011). Reciprocal sign epistasis between frequently experimentally evolved adaptive mutations causes a rugged fitness landscape. PLoS Genetics, 7(4), e1002056.

    Article  Google Scholar 

  • Lake, M. W., & Venti, J. (2009). Quantitative analysis of macroevolutionary patterning in technological evolution: bicycle design from 1800 to 2000. In S. J. Shennan (Ed.), Pattern and process in cultural evolution (pp. 147–162). Berkeley: University of California Press.

    Google Scholar 

  • Laland, K. N. (2004). Social learning strategies. Learning & Behavior, 32, 4–14.

    Article  Google Scholar 

  • Laland, K. N., Atton, N., & Webster, M. M. (2011). From fish to fashion: experimental and theoretical insights into the evolution of culture. Philosophical Transactions of the Royal Society B, 366, 958–968.

    Article  Google Scholar 

  • Leonard, R. D. (2001). Evolutionary archaeology. In I. Hodder (Ed.), Archaeological theory today (pp. 65–97). Cambridge: Polity.

    Google Scholar 

  • Leonard, R. D., & Jones, G. T. (1987). Elements of an inclusive evolutionary model for archaeology. Journal of Anthropological Archaeology, 6, 199–219.

    Article  Google Scholar 

  • Lipo, C. P. (2006). The resolution of cultural phylogenies using graphs. In C. P. Lipo, M. J. O’Brien, M. Collard, & S. Shennan (Eds.), Mapping our ancestors: phylogenetic methods in anthropology and prehistory (pp. 89–107). New York: Aldine.

    Google Scholar 

  • Lipo, C. P., O’Brien, M. J., Collard, M., & Shennan, S. J. (Eds.). (2006). Mapping our ancestors: phylogenetic approaches in anthropology and prehistory. New York: Aldine.

    Google Scholar 

  • Lipo, C. P., Dunnell, R. C., O’Brien, M. J., Harper, V., & Dudgeon, J. (2012). Beveled projectile points and ballistics technology. American Antiquity, 77, 774–788.

    Article  Google Scholar 

  • Lobo, J., & Macready, W. G. (1999). Landscapes: a natural extension of search theory. Santa Fe Institute Working Paper 99-05-037E. Santa Fe, NM.

  • Lothrop, J. C., Newby, P. E., Spiess, A. E., & Bradley, J. (2011). Paleoindians and the Younger Dryas in the New England–Maritimes region. Quaternary International, 242, 546–569.

    Article  Google Scholar 

  • Lycett, S. J. (2008). Acheulean variation and selection: does handaxe symmetry fit neutral expectations? Journal of Archaeological Science, 35, 2640–2648.

    Article  Google Scholar 

  • Lycett, S. J. (2009a). Understanding ancient hominin dispersals using artefactual data: a phylogeographic analysis of Acheulean handaxes. PLoS ONE, 4(10), e7404–1–6.

    Article  Google Scholar 

  • Lycett, S. J. (2009b). Are Victoria West cores ‘proto-Levallois’? A phylogenetic assessment. Journal of Human Evolution, 56, 175–191.

    Article  Google Scholar 

  • Lycett, S. J. (2010). The importance of history in definitions of ‘culture’: implications from phylogenetic approaches to the study of social learning in chimpanzees. Learning & Behavior, 38, 252–264.

    Article  Google Scholar 

  • Lycett, S. J. (2014). Dynamics of cultural transmission in Native Americans of the High Great Plains. PLoS ONE, 9(11), e112244.

    Article  Google Scholar 

  • Lycett, S. J. (2015). Cultural evolutionary approaches to artifact variation over time and space: basis, progress, and prospects. Journal of Archaeological Science, 56, 21–31.

  • Lycett, S. J., & Eren, M. I. (2013). Levallois lessons: the challenge of integrating mathematical models, experiments and the archaeological record. World Archaeology, 45, 519–538.

    Article  Google Scholar 

  • Lycett, S. J., & von Cramon-Taubadel, N. (2008). Acheulean variability and hominin dispersals: a model-bound approach. Journal of Archaeological Science, 35, 553–562.

    Article  Google Scholar 

  • Lycett, S. J., & von Cramon-Taubadel, N. (2013). A 3D morphometric analysis of surface geometry in Levallois cores: patterns of stability and variability across regions and their implications. Journal of Archaeological Science, 40, 1508–1517.

    Article  Google Scholar 

  • Lycett, S. J., & von Cramon-Taubadel, N. (2015). Toward a “quantitative genetic” approach to lithic variation. Journal of Archaeological Method and Theory, 22: 646–675.

  • Lycett, S. J., von Cramon-Taubadel, N., & Gowlett, J. A. (2010). A comparative 3D geometric morphometric analysis of Victoria west cores: implications for the origins of Levallois technology. Journal of Archaeological Science, 37, 1110–1117.

    Article  Google Scholar 

  • Lyman, R. L., & O’Brien, M. J. (1998). The goals of evolutionary archaeology: history and explanation. Current Anthropology, 39, 615–652.

    Article  Google Scholar 

  • Lyman, R. L., & O’Brien, M. J. (2001). Misconceptions of evolutionary archaeology: confusing macroevolution and microevolution. Current Anthropology, 42, 408–409.

    Article  Google Scholar 

  • Lyman, R. L., & O’Brien, M. J. (2002). Classification. In J. P. Hart & J. E. Terrell (Eds.), Darwin and archaeology: a handbook of key concepts (pp. 69–88). Westport, CT: Bergin & Garvey.

    Google Scholar 

  • Lyman, R. L., & O’Brien, M. J. (2003). Cultural traits: units of analysis in early twentieth-century anthropology. Journal of Anthropological Research, 59, 225–250.

    Article  Google Scholar 

  • Lyman, R. L., & O’Brien, M. J. (2005). Measuring time with artifacts: a history of methods in American archaeology. Lincoln: University of Nebraska Press.

    Google Scholar 

  • Lyman, R. L., & O’Brien, M. J. (2006). Seriation and cladistics: the difference between anagenetic and cladogenetic evolution. In C. P. Lipo, M. J. O’Brien, S. J. Shennan, & M. Collard (Eds.), Mapping our ancestors: phylogenetic methods in anthropology and prehistory (pp. 65–88). New York: Aldine.

    Google Scholar 

  • Lyman, R. L., O’Brien, M. J., & Hayes, V. (1998). A mechanical and functional study of bone rods from the Richey–Roberts Clovis Cache, Washington, U.S.A. Journal of Archaeological Science, 25, 887–906.

    Article  Google Scholar 

  • Lyman, R. L., VanPool, T. L., & O’Brien, M. J. (2008). Variation in North American dart points and arrow points when one or both are present. Journal of Archaeological Science, 35, 2805–2812.

    Article  Google Scholar 

  • Lyman, R. L., VanPool, T. L., & O’Brien, M. J. (2009). The diversity of North American projectile-point types, before and after the bow and arrow. Journal of Anthropological Archaeology, 28, 1–13.

    Article  Google Scholar 

  • MacDonald, G. F. (1968). Debert: a Palaeo-Indian site in central Nova Scotia. Ottawa: National Museum of Man.

    Google Scholar 

  • Mace, R., & Pagel, M. (1994). The comparative method in anthropology. Current Anthropology, 35, 549–564.

    Article  Google Scholar 

  • Mace, R., Holden, C., & Shennan, S. J. (Eds.). (2005). The evolution of cultural diversity: a phylogenetic approach. London: University College London Press.

    Google Scholar 

  • Maddison, W. P., Donoghue, M. J., & Maddison, D. R. (1984). Outgroup analysis and parsimony. Systematic Zoology, 33, 83–103.

    Article  Google Scholar 

  • Madsen, M., Lipo, C., & Cannon, M. (1999). Fitness and reproductive trade-offs in uncertain environments: explaining the evolution of cultural elaboration. Journal of Anthropological Archaeology, 18, 251–281.

    Article  Google Scholar 

  • Mallet, J. (2005). Hybridization as an invasion of the gene. Trends in Ecology and Evolution, 20, 229–237.

    Article  Google Scholar 

  • Mason, R. J. (1962). The Paleo-Indian tradition in eastern North America. Current Anthropology, 3, 227–278.

    Article  Google Scholar 

  • McGhee, G. R. (2011). Convergent evolution: limited forms most beautiful. Cambridge, MA: MIT Press.

    Book  Google Scholar 

  • McNett, C. W., Jr. (1979). The cross-cultural method in archaeology. In M. B. Schiffer (Ed.), Advances in archaeological method and theory (vol. 2) (pp. 39–76). New York: Academic Press.

    Google Scholar 

  • Meltzer, D. J. (2009). First peoples in a new world: colonizing Ice Age America. Berkeley: University of California Press.

    Google Scholar 

  • Mesoudi, A. (2008). An experimental simulation of the “copy-successful-individuals” cultural learning strategy: adaptive landscapes, producer–scrounger dynamics, and informational access costs. Evolution and Human Behavior, 29, 350–363.

    Article  Google Scholar 

  • Mesoudi, A. (2010). The experimental study of cultural innovation. In M. J. O’Brien & S. J. Shennan (Eds.), Innovation in cultural systems: contributions from evolutionary anthropology (pp. 175–191). Cambridge, MA: MIT Press.

    Google Scholar 

  • Mesoudi, A. (2011a). Cultural evolution: how Darwinian evolutionary theory can explain human culture and synthesize the social sciences. Chicago: University of Chicago Press.

    Book  Google Scholar 

  • Mesoudi, A. (2011b). An experimental comparison of human social learning strategies: payoff-biased social learning is adaptive but under-used. Evolution and Human Behavior, 32, 334–342.

    Article  Google Scholar 

  • Mesoudi, A. (2014). Experimental studies of modern human social and individual learning in an archaeological context: people behave adaptively, but within limits. In T. Akazawa, N. Ogilhara, H. C. Tanabe, H. Terashima, T. Akazawa, N. Ogilhara, H. C. Tanabe, & H. Terashima (Eds.), Dynamics of learning in Neanderthals and modern humans (Vol. 2): cognitive and physical perspectives (pp. 65–76). Tokyo: Springer.

    Chapter  Google Scholar 

  • Mesoudi, A., & Lycett, S. J. (2009). Random copying, frequency-dependent copying and culture change. Evolution and Human Behavior, 30, 41–48.

    Article  Google Scholar 

  • Mesoudi, A., & O’Brien, M. J. (2008a). The cultural transmission of Great Basin projectile-point technology I: an experimental simulation. American Antiquity, 73, 3–28.

    Google Scholar 

  • Mesoudi, A., & O’Brien, M. J. (2008b). The cultural transmission of Great Basin projectile-point technology II: an agent-based computer simulation. American Antiquity, 73, 627–724.

    Google Scholar 

  • Mesoudi, A., Whiten, A., & Laland, K. N. (2006). Towards a unified science of cultural evolution. Behavioral and Brain Sciences, 29, 329–383.

    Google Scholar 

  • Miller, G. L. (2014). Lithic microwear analysis as a means to infer production of perishable technology: a case from the Great Lakes. Journal of Archaeological Science, 49, 292–301.

    Article  Google Scholar 

  • Morrow, J. E. (2015). Clovis-era point production in the Midcontinent. In A. M. Smallwood & T. A. Jennings (Eds.), Clovis: on the edge of a new understanding (pp. 83–107). College Station: Texas A&M University Press.

    Google Scholar 

  • Morrow, J. E., & Morrow, T. A. (1999). Geographic variation in fluted projectile points: a hemispheric perspective. American Antiquity, 64, 215–231.

    Article  Google Scholar 

  • Naroll, R. (1970). Galton’s problem. In R. Naroll & R. Cohen (Eds.), A handbook of method in cultural anthropology (pp. 974–989). New York: Columbia University Press.

    Google Scholar 

  • Nixon, K. C., & Carpenter, J. M. (2012). More on homology. Cladistics, 28, 225–226.

    Article  Google Scholar 

  • Nolan, K. C., & Cook, R. A. (2011). A critique of late prehistoric systematics in the middle Ohio River valley. North American Archaeologist, 32, 293–325.

    Article  Google Scholar 

  • Nunn, C. L., Borgerhoff Mulder, M., & Langley, S. (2006). Comparative methods for studying cultural trait evolution: a simulation study. Cross-Cultural Research, 40, 177–209.

    Article  Google Scholar 

  • Nunn, C. L., Arnold, C., Matthews, L., & Borgerhoff Mulder, M. (2010). Simulating trait evolution for cross-cultural comparison. Philosophical Transactions of the Royal Society B, 365, 3807–3819.

    Article  Google Scholar 

  • O’Brien, M. J. (1987). Sedentism, population growth, and resource selection in the Woodland Midwest: a review of coevolutionary developments. Current Anthropology, 28, 177–197.

    Article  Google Scholar 

  • O’Brien, M. J., & Bentley, R. A. (2011). Stimulated variation and cascades: two processes in the evolution of complex technological systems. Journal of Archaeological Method and Theory, 18, 309–335.

    Article  Google Scholar 

  • O’Brien, M. J., & Holland, T. D. (1992). The role of adaptation in archaeological explanation. American Antiquity, 57, 36–59.

    Article  Google Scholar 

  • O’Brien, M. J., & Holland, T. D. (1995). Behavioral archaeology and the extended phenotype. In J. M. Skibo, W. H. Walker, & A. E. Nielsen (Eds.), Expanding archaeology (pp. 143–161). Salt Lake City: University of Utah Press.

    Google Scholar 

  • O’Brien, M. J., & Lyman, R. L. (1999). Seriation, stratigraphy, and index fossils: the backbone of archaeological dating. New York: Kluwer Academic/Plenum.

    Google Scholar 

  • O’Brien, M. J., & Lyman, R. L. (2000). Applying evolutionary archaeology: a systematic approach. New York: Kluwer Academic/Plenum.

    Google Scholar 

  • O’Brien, M. J., & Lyman, R. L. (2002a). Evolutionary archaeology: current status and future prospects. Evolutionary Anthropology, 11, 26–36.

    Article  Google Scholar 

  • O’Brien, M. J., & Lyman, R. L. (2002b). The epistemological nature of archaeological units. Anthropological Theory, 2, 37–56.

    Article  Google Scholar 

  • O’Brien, M. J., & Lyman, R. L. (2003). Style, function, transmission: an introduction. In M. J. O’Brien & R. L. Lyman (Eds.), Style, function, transmission: evolutionary archaeological perspectives (pp. 1–32). Salt Lake City: University of Utah Press.

    Google Scholar 

  • O’Brien, M. J., & Shennan, S. J. (Eds.). (2010). Innovation in cultural systems: contributions from evolutionary anthropology. Cambridge, MA: MIT Press.

    Google Scholar 

  • O’Brien, M. J., Holland, T. D., Hoard, R. J., & Fox, G. L. (1994). Evolutionary implications of design and performance characteristics of prehistoric pottery. Journal of Archaeological Method and Theory, 1, 259–304.

    Article  Google Scholar 

  • O’Brien, M. J., Darwent, J., & Lyman, R. L. (2001). Cladistics is useful for reconstructing archaeological phylogenies: Palaeoindian points from the southeastern United States. Journal of Archaeological Science, 28, 1115–1136.

    Article  Google Scholar 

  • O’Brien, M. J., Lyman, R. L., Saab, Y., Saab, E., Darwent, J., & Glover, D. S. (2002). Two issues in archaeological phylogenetics: taxon construction and outgroup selection. Journal of Theoretical Biology, 215, 133–150.

    Article  Google Scholar 

  • O’Brien, M. J., Lyman, R. L., & Leonard, R. D. (2003). What is evolution? A reply to Bamforth. American Antiquity, 68, 573–580.

    Article  Google Scholar 

  • O’Brien, M. J., Lyman, R. L., Mesoudi, A., & VanPool, T. L. (2010). Cultural traits as units of analysis. Philosophical Transactions of the Royal Society B, 365, 3797–3806.

    Article  Google Scholar 

  • O’Brien, M. J., Buchanan, B., Collard, M., & Boulanger, M. T. (2012). Cultural cladistics and the early history of North America. In P. Pontarotti (Ed.), Evolutionary biology: mechanisms and trends (pp. 23–42). New York: Springer.

    Chapter  Google Scholar 

  • O’Brien, M. J., Collard, M., Buchanan, B., & Boulanger, M. T. (2013). Trees, thickets, or something in between? Recent theoretical and empirical work in cultural phylogeny. Israel Journal of Ecology and Evolution, 59(2), 49–61.

    Article  Google Scholar 

  • O’Brien, M. J., Boulanger, M. T., Buchanan, B., Collard, M., Lyman, R. L., & Darwent, J. (2014). Innovation and cultural transmission in the American Paleolithic: phylogenetic analysis of eastern Paleoindian projectile-point classes. Journal of Anthropological Archaeology, 34, 100–119.

    Article  Google Scholar 

  • O’Brien, M. J., Boulanger, M. T., Lyman, R. L., & Buchanan, B. (2015a). Phylogenetic systematics. In J. A. Barceló & I. Bogdanovic (Eds.), Mathematics in archaeology (pp. 232–246). Boca Raton, FL: CRC Press.

    Google Scholar 

  • O’Brien, M. J., Buchanan, B., Boulanger, M. T., Mesoudi, A., Collard, M., Eren, M. I., Bentley, R. A., & Lyman, R. L. (2015b). Transmission of cultural variants in the North American Paleolithic. In K. Aoki & A. Mesoudi (Eds.), Learning strategies and cultural evolution during the Paleolithic (pp. 121–143). New York: Springer.

    Google Scholar 

  • Olausson, D. (2008). Does practice make perfect? Craft expertise as a factor in aggrandizer strategies. Journal of Archaeological Method and Theory, 15, 28–50.

    Article  Google Scholar 

  • Östborn, P., & Gerding, H. (2014). Network analysis of archaeological data: a systematic approach. Journal of Archaeological Science, 46, 75–88.

    Article  Google Scholar 

  • Östborn, P., & Gerding, H. (2015). The diffusion of fired bricks in Hellenistic Europe: a similarity network analysis. Journal of Archaeological Method and Theory, 22, 306–344.

  • Page, R. D. M. (Ed.). (2003). Tangled trees: phylogeny, cospeciation, and coevolution. Chicago: University of Chicago Press.

    Google Scholar 

  • Pagel, M. (1999). Inferring the historical patterns of biological evolution. Nature, 401, 877–884.

    Article  Google Scholar 

  • Patten, B. (2005). Peoples of the flute: a study in anthropolithic forensics. Lakewood, CO: Stone Dagger Publications.

    Google Scholar 

  • Patterson, C. (1988). Homology in classical and molecular biology. Molecular Biology and Evolution, 5, 603–625.

    Google Scholar 

  • Pigeot, N. (1990). Technical and social actors: flint knapping specialists and apprentices at Magdalenian Etiolles. Archaeological Review from Cambridge, 9, 126–141.

    Google Scholar 

  • Prasciunas, M. M. (2011). Mapping Clovis: projectile points, behavior, and bias. American Antiquity, 76, 107–126.

    Article  Google Scholar 

  • Prasciunas, M. M., & Surovell, T. A. (2015). Reevaluating the duration of Clovis: the problem of non-representative radiocarbon. In A. M. Smallwood & T. A. Jennings (Eds.), Clovis: on the edge of a new understanding (pp. 21–35). College Station: Texas A&M University Press.

    Google Scholar 

  • Premo, L. S., & Scholnick, J. B. (2011). The spatial scale of social learning affects cultural diversity. American Antiquity, 76, 163–176.

    Article  Google Scholar 

  • Prentiss, A. M., Chatters, J. C., Walsh, M. W., & Skelton, R. R. (2013). Cultural macroevolution in the Pacific northwest: a phylogenetic test of the diversification and decimation model. Journal of Archaeological Science, 41, 29–43.

    Article  Google Scholar 

  • Prufer, O. H., & Baby, R. S. (1963). Palaeo-Indians of Ohio. Columbus: Ohio Historical Society.

    Google Scholar 

  • Ramenofsky, A. F. (1998). Evolutionary theory and the Native American record of artifact replacement. In J. G. Cusick (Ed.), Studies in culture contact: interaction, culture change, and archaeology. Center for Archaeological Investigations Occasional Paper 25, pp. 77–101. Carbondale: Southern Illinois University Press.

    Google Scholar 

  • Rendell, L., Boyd, R., Cownden, D., Enquist, M., Eriksson, K., Feldman, M. W., Ghirlanda, S., Lillicrap, T., & Laland, K. N. (2010). Why copy others? Insights from the Social Learning Strategies Tournament. Science, 328, 208–213.

    Article  Google Scholar 

  • Rendell, L., Boyd, R., Enquist, M., Feldman, M. W., Fogarty, L., & Laland, K. N. (2011). How copying affects the amount, evenness and persistence of cultural knowledge: insights from the Social Learning Strategies Tournament. Philosophical Transactions of the Royal Society B, 366, 1118–1128.

    Article  Google Scholar 

  • Revell, L. J., Harmon, L. J., & Collar, D. C. (2008). Phylogenetic signal, evolutionary process, and rate. Systematic Biology, 57, 591–601.

    Article  Google Scholar 

  • Rexová, K., Frynta, D., & Zrzávy, J. (2003). Cladistic analysis of languages: Indo-European classification based on lexicostatistical data. Cladistics, 19, 120–127.

    Article  Google Scholar 

  • Reyes-Garcia, V., Molina, J. L., Broesch, J., Calvet, L., Huanca, T., Saus, J., Tanner, S., Leonard, W. R., & McDade, T. W. (2008). Do the aged and knowledgeable men enjoy more prestige? A test of predictions from the prestige-bias model of cultural transmission. Evolution and Human Behavior, 29, 275–281.

    Article  Google Scholar 

  • Rhymer, J., & Simberloff, D. (1996). Extinction by hybridization and introgression. Annual Review of Ecology and Systematics, 27, 83–109.

    Article  Google Scholar 

  • Richerson, P. J., & Boyd, R. (2000). Climate, culture, and the evolution of cognition. In C. Heyes & L. Huber (Eds.), Evolution of cognition (pp. 329–346). Cambridge, MA: MIT Press.

    Google Scholar 

  • Richerson, P. J., & Boyd, R. (2005). Not by genes alone. Chicago: University of Chicago Press.

    Google Scholar 

  • Riede, F. (2011). Adaptation and niche construction in human prehistory: a case study from the southern Scandinavian Late Glacial. Philosophical Transactions of the Royal Society B, 366, 793–808.

    Article  Google Scholar 

  • Ritchie, W. A. (1961). New York projectile points: a typology and nomenclature. New York State Museum, Bulletin 384. Albany.

  • Roosa, W. B. (1965). Some Great Lakes fluted point types. Michigan Archaeologist, 11(3–4), 89–102.

    Google Scholar 

  • Rosewich, U. L., & Kistler, H. C. (2000). Role of horizontal gene transfer in the evolution of fungi. Annual Review of Phytopathology, 38, 325–363.

    Article  Google Scholar 

  • Ross, R. M., Grenhill, S. J., & Atkinson, Q. (2013). Population structure and cultural geography of a folktale in Europe. Proceedings of the Royal Society B, 280, 20123065.

    Article  Google Scholar 

  • Roy, K., Valentine, J. W., Jablonski, D., & Kidwell, S. M. (1996). Scales of climatic variability and time averaging in Pleistocene biotas: implications for ecology and evolution. Trends in Ecology and Evolution, 11, 458–463.

    Article  Google Scholar 

  • Sanchez, G., Holliday, V. T., Gaines, E. P., Arroyo-Cabrales, J., Martínez-Tagüeña, N., Kowler, A., Lange, T., Hodgins, G. W. L., Mentzer, S. M., & Sanchez-Morales, I. (2014). Human (Clovis)–gomphothere (Cuvieronius sp.) association ∼13,390 calibrated yBP in Sonora, Mexico. Proceedings of the National Academy of Sciences, 111, 10972–10977.

    Article  Google Scholar 

  • Sanderson, M. J., & Donoghue, M. J. (1989). Patterns of variation in levels of homoplasy. Evolution, 43, 1781–1795.

    Article  Google Scholar 

  • Sanderson, M. J., & Hufford, L. (Eds.). (1996). Homoplasy: the recurrence of similarity in evolution. New York: Academic Press.

    Google Scholar 

  • Schiffer, M. B., & Skibo, J. M. (1989). A provisional theory of ceramic abrasion. American Anthropologist, 91, 101–115.

    Article  Google Scholar 

  • Schillinger, K., Mesoudi, A., & Lycett, S. J. (2014a). Copying error and the cultural evolution of ‘additive’ versus ‘reductive’ material traditions: an experimental assessment. American Antiquity, 79, 128–143.

    Article  Google Scholar 

  • Schillinger, K., Mesoudi, A., & Lycett, S. J. (2014b). Considering the role of time budgets on copy-error rates in material culture traditions: an experimental assessment. PLoS ONE, 9(5), e97157.

    Article  Google Scholar 

  • Schumpeter, J. A. (1942). Capitalism, socialism, and democracy. New York: Harper.

    Google Scholar 

  • Shaw, A. B. (1969). Adam and Eve, paleontology, and the non-objective arts. Journal of Paleontology, 43, 1085–1098.

    Google Scholar 

  • Shennan, S. J. (2002). Genes, memes and human history. London: Thames and Hudson.

    Google Scholar 

  • Shennan, S. J. (Ed.). (2009). Pattern and process in cultural evolution. Berkeley: University of California Press.

    Google Scholar 

  • Sholts, S. B., Stanford, D. J., Flores, L. M., & Wärmländer, S. K. T. S. (2012). Flake scar patterns of Clovis points analyzed with a new digital morphometrics approach: evidence for direct transmission of technological knowledge across early North America. Journal of Archaeological Science, 39, 3018–3026.

    Article  Google Scholar 

  • Shott, M. J., & Trail, B. W. (2010). Exploring new approaches to lithic analysis: laser scanning and geometric morphometrics. Lithic Technology, 35, 195–220.

    Article  Google Scholar 

  • Simmons, M. P. (2012). Misleading results of likelihood-based phylogenetic analyses in the presence of missing data. Cladistics, 28, 208–222.

    Article  Google Scholar 

  • Simpson, G. G. (1961). Principles of animal taxonomy. New York: Columbia University Press.

    Google Scholar 

  • Skibo, J. M., Schiffer, M. B., & Reid, K. C. (1989). Organic-tempered pottery: an experimental study. American Antiquity, 54, 122–146.

    Article  Google Scholar 

  • Slice, D. E. (2007). Geometric morphometrics. Annual Review of Anthropology, 36, 261–281.

    Article  Google Scholar 

  • Smallwood, A. M. (2012). Clovis technology and settlement in the American Southeast: using biface analysis to evaluate dispersal models. American Antiquity, 77, 689–713.

    Article  Google Scholar 

  • Smallwood, A. M. (2015). Building experimental use-wear analogues for Clovis biface functions. Archaeological and Anthropological Sciences, 7, 13–26.

    Article  Google Scholar 

  • Smallwood, A. M., & Jennings, T. A. (Eds.). (2015). Clovis: on the edge of a new understanding. College Station: Texas A&M University Press.

    Google Scholar 

  • Smith, H. L., Smallwood, A. M., & DeWitt, T. J. (2015). Defining the normative range of Clovis fluted point shape using geographic models of geometric morphometric variation. In A. M. Smallwood & T. A. Jennings (Eds.), Clovis: on the edge of a new understanding (pp. 161–180). College Station: Texas A&M University Press.

    Google Scholar 

  • Stark, M. T., Bowser, B. J., & Horne, L. (Eds.). (2008). Cultural transmission and material culture: breaking down boundaries. Tucson: University of Arizona Press.

    Google Scholar 

  • Steffensen, J. P., Andersen, K. K., Bigler, M., Clausen, H. B., Dahl-Jensen, D., Fuscher, H., Goto-Azuma, K., Hansson, M., Johnsen, S. J., Jouzel, J., Masson-Delmotte, V., Popp, T., Rasmussen, S. O., Röthlisberger, R., Ruth, U., Stauffer, B., Siggaard-Andersen, M.-L., Sveinbjörnsdóttir, A. E., Svensson, A., & White, J. W. C. (2008). High-resolution Greenland ice core data show abrupt climate change happens in few years. Science, 321, 680–684.

    Article  Google Scholar 

  • Steward, J. H. (1954). Types of types. American Anthropologist, 56, 54–57.

    Google Scholar 

  • Stuart, T. E., & Podolny, J. M. (1996). Local search and the evolution of technological capabilities. Strategic Management Journal, 17, 21–23.

    Article  Google Scholar 

  • Swofford, D. (1991). When are phylogeny estimates from morphological and molecular data incongruent? In M. M. Miyamoto & J. Cracraft (Eds.), Phylogenetic analysis of DNA sequences (pp. 295–333). New York: Oxford University Press.

    Google Scholar 

  • Swofford, D. (1998). PAUP•: phylogenetic analysis using parsimony (*and other methods) (version 4). Sunderland, MA: Sinauer.

    Google Scholar 

  • Tankersley, K. B. (1989). Late Pleistocene lithic exploitation and human settlement in the midwestern United States. PhD dissertation, Department of Anthropology, Indiana University. Bloomington.

  • Taylor, K. C., Lamorey, G. W., Doyle, G. A., Alley, R. B., Grootes, P. M., Mayewski, P. A., White, J. W. C., & Barlow, L. K. (1993). The ‘flickering switch’ of late Pleistocene climate change. Nature, 361, 432–436.

    Article  Google Scholar 

  • Tehrani, J. J. (2013). The phylogeny of Little Red Riding Hood. PLoS ONE, 8(11), e78871.

    Article  Google Scholar 

  • Tehrani, J. J., & Collard, M. (2002). Investigating cultural evolution through biological phylogenetic analyses of Turkmen textiles. Journal of Anthropological Archaeology, 21, 443–463.

    Article  Google Scholar 

  • Tehrani, J. J., Collard, M., & Shennan, S. J. (2010). The cophylogeny of populations and cultures: reconstructing the evolution of Iranian tribal craft traditions using trees and jungles. Philosophical Transactions of the Royal Society B, 365, 3865–3874.

    Article  Google Scholar 

  • Thulman, D. K. (2012). Discriminating Paleoindian point types from Florida using landmark geometric morphometrics. Journal of Archaeological Science, 39, 1599–1607.

    Article  Google Scholar 

  • Thulman, D. K. (2013). The role of nondeclarative memory systems in the inference of long-term population continuity. Journal of Archaeological Method and Theory, 21, 724–749.

    Article  Google Scholar 

  • Toelch, U., van Delft, M. J., Bruce, M. J., Donders, R., Meeus, M. T. H., & Reader, S. M. (2009). Decreased environmental variability induces a bias for social information use in humans. Evolution and Human Behavior, 30, 32–40.

    Article  Google Scholar 

  • Tomasello, M., Kruger, A. C., & Ratner, H. H. (1993). Cultural learning. Behavioral and Brain Sciences, 16, 495–511.

    Article  Google Scholar 

  • Tostevin, G. (2012). Seeing lithics: a middle range theory for testing for cultural transmission in the Pleistocene. Oxford: Oxbow Books.

    Google Scholar 

  • Waguespack, N. M. (2007). Why we’re still arguing about the Pleistocene occupation of the Americas. Evolutionary Anthropology, 16, 63–74.

    Article  Google Scholar 

  • Waguespack, N. M., Surovell, T. A., Denoyer, A., Dallow, A., Savage, A., Hyneman, J., & Tapster, D. (2009). Making a point: wood versus stone-tipped projectiles. Antiquity, 83, 786–800.

    Article  Google Scholar 

  • Walker, M., Johnsen, S., Rasmussen, S. O., Popp, T., Steffensen, J.-P., Gibbard, P., Hoek, W., Lowe, J., Andrews, J., Bjorck, S., Cwynar, L. C., Hughen, K., Kershaw, P., Kromer, B., Litt, T., Lowe, D. J., Nakagawa, T., Newnham, R., & Schwander, J. (2009). Formal definition and dating of the GSSP (Global Stratotype Section and Point) for the base of the Holocene using the Greenland NGRIP ice core, and selected auxiliary records. Journal of Quaternary Science, 24, 3–17.

    Article  Google Scholar 

  • Wang, W., Lycett, S. J., von Cramon-Taubadel, N., Jin, J. J. H., & Bae, C. J. (2012). Comparison of handaxes from Bose Basin (China) and the western Acheulean indicates convergence of form, not cognitive differences. PLoS ONE, 7(4), e35804.

    Article  Google Scholar 

  • Waters, M. R., & Stafford, T. W., Jr. (2007). Redefining the age of Clovis: implications for the peopling of the Americas. Science, 315, 1122–1126.

    Article  Google Scholar 

  • Whiten, A. (2005). The second inheritance system of chimpanzees and humans. Nature, 437, 52–55.

    Article  Google Scholar 

  • Whiten, A., Hinde, R. A., Laland, K. N., & Stringer, C. B. (2011). Culture evolves. Philosophical Transactions of the Royal Society B, 366, 938–948.

    Article  Google Scholar 

  • Whitt, E. E. (2010). Extraordinary fluted points of the Tennessee Valley region. Florence, AL: Printers and Stationers.

    Google Scholar 

  • Whittaker, J. C. (1994). Flintknapping: making and understanding stone tools. Austin: University of Texas Press.

    Google Scholar 

  • Whittaker, J. C. (2004). American flintknappers: Stone age art in the age of computers. Austin: University of Texas Press.

    Google Scholar 

  • Williams, P. A. (2002). Of replicators and selectors. Quarterly Review of Biology, 77, 302–306.

    Article  Google Scholar 

  • Wormington, H. M. (1957). Ancient man in North America (4th ed.). Denver Museum of Natural History Popular Series, No. 4. Denver.

  • Wright, S. (1932). The roles of mutation, inbreeding, crossbreeding and selection in evolution. In D. F. Jones (Ed.), Proceedings of The Sixth Congress on Genetics (vol. 1) (pp. 356–366). New York: Brooklyn Botanic Garden.

    Google Scholar 

  • Wright, S. (1988). Surfaces of selective value revisited. American Naturalist, 131, 115–123.

    Article  Google Scholar 

  • Zelditch, M. L., Swiderski, D. L., Sheets, H. D., & Fink, W. L. (2004). Geometric morphometrics for biologists: a primer. Amsterdam: Elsevier.

    Google Scholar 

Download references

Acknowledgments

We thank Jim Skibo, Cathy Cameron, and Ashley Smallwood, along with two anonymous reviewers, for numerous suggestions for improving the original manuscript. MIE is supported by a postdoctoral fellowship from the University of Missouri.

Compliance with Ethical Standards

Conflict of Interest

We have no disclosures of potential conflicts of interest.

Ethics Approval and Consent to Participate

Our research did not involve human participants and/or animals. Because our research involved no human participants, informed consent is not applicable.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michael J. O’Brien.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

O’Brien, M.J., Boulanger, M.T., Buchanan, B. et al. Design Space and Cultural Transmission: Case Studies from Paleoindian Eastern North America. J Archaeol Method Theory 23, 692–740 (2016). https://doi.org/10.1007/s10816-015-9258-7

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10816-015-9258-7

Keywords

Navigation