Sie sind auf Seite 1von 32

Paleoindian settlement of the high-altitude Peruvian Andes

Kurt Rademaker et al.


Science 346, 466 (2014);
DOI: 10.1126/science.1258260

This copy is for your personal, non-commercial use only.

If you wish to distribute this article to others, you can order high-quality copies for your
colleagues, clients, or customers by clicking here.
Permission to republish or repurpose articles or portions of articles can be obtained by
following the guidelines here.

The following resources related to this article are available online at


www.sciencemag.org (this information is current as of October 24, 2014 ):

Downloaded from www.sciencemag.org on October 24, 2014


Updated information and services, including high-resolution figures, can be found in the online
version of this article at:
http://www.sciencemag.org/content/346/6208/466.full.html
Supporting Online Material can be found at:
http://www.sciencemag.org/content/suppl/2014/10/22/346.6208.466.DC1.html
A list of selected additional articles on the Science Web sites related to this article can be
found at:
http://www.sciencemag.org/content/346/6208/466.full.html#related
This article cites 58 articles, 9 of which can be accessed free:
http://www.sciencemag.org/content/346/6208/466.full.html#ref-list-1
This article appears in the following subject collections:
Anthropology
http://www.sciencemag.org/cgi/collection/anthro

Science (print ISSN 0036-8075; online ISSN 1095-9203) is published weekly, except the last week in December, by the
American Association for the Advancement of Science, 1200 New York Avenue NW, Washington, DC 20005. Copyright
2014 by the American Association for the Advancement of Science; all rights reserved. The title Science is a
registered trademark of AAAS.
R ES E A RC H | R E PO R TS

interspecific hybridization; yet, such reproduc- 26. R. D. Holt, Theor. Popul. Biol. 12, 19729 (1977). T.S.C., and J.B.L. designed the study; Y.E.S., T.S.C., P.A.H., L.J.R,
tive character displacement (4) seems an unlike- 27. G. A. Polis, C. A. Myers, R. D. Holt, Annu. Rev. Ecol. Syst. 20, and R.G.R. collected the data; Y.E.S., T.S.C., and P.A.H. analyzed
297330 (1989). the data; all authors contributed to the manuscript. Data are
ly explanation for our results, as A. carolinensis 28. R. Shine, Evol. Appl. 5, 107116 (2012). accessioned on datadryad.org: doi:10.5061/dryad.96g44.
and A. sagrei already differ markedly in species-
recognition characteristics, males of both species ACKN OWLED GMEN TS
SUPPLEMENTARY MATERIALS
nearly exclusively ignore heterospecific females in We thank A. Kamath, C. Gilman, A. Algar, J. Allen, E. Boates,
www.sciencemag.org/content/346/6208/463/suppl/DC1
staged encounters (25), and the species have never A. Echternacht, A. Harrison, H. Lyons-Galante, T. Max, J. McCrae,
Materials and Methods
J. Newman, J. Rifkin, M. Stimola, P. VanMiddlesworth, K. Winchell,
been reported to successfully produce hybrids. We C. Wiench, K. Wollenberg, and three reviewers; M. Legare and
Supplementary Acknowledgments
note, finally, that other mutually negative inter- Figs. S1 and S2
J. Lyon (Merritt Island National Wildlife Refuge), J. Stiner and
Tables S1 to S7
actions such as apparent competition (26) and in- C. Carter (Canaveral National Seashore); and Harvard University,
References (2945)
traguild predation (27) could also produce Museum of Comparative Zoology, University of Massachusetts
Boston, University of Tennessee Knoxville, University of Tampa, 5 June 2014; accepted 15 September 2014
divergence among overlapping species. These re- NSF (DEB-1110521), and NIH (P30GM103324) for funding. Y.E.S., 10.1126/science.1257008
main to be explored in this system, though some
evidence exists for at least the latter (17).
Here, we have provided evidence from a repli-
cated, natural system to support the long-held NEW WORLD ARCHAEOLOGY
idea (4) that interspecific interactions between
closely related species are an important force for
evolutionary diversification (2). Moreover, we show
that evolutionary hypotheses such as character
Paleoindian settlement of the
displacement can be rigorously tested in real
time following human-caused environmental high-altitude Peruvian Andes
change. Our results also demonstrate that native
species may be able to respond evolutionarily to Kurt Rademaker,1,2,8* Gregory Hodgins,3 Katherine Moore,4 Sonia Zarrillo,5
strong selective forces wrought by invaders. The Christopher Miller,6,7 Gordon R. M. Bromley,8 Peter Leach,9 David A. Reid,10
extent to which the costs of invasions can be mit- Willy Ypez lvarez,11 Daniel H. Sandweiss1,8
igated by evolutionary response remains to be
determined (28), but studies such as this demon- Study of human adaptation to extreme environments is important for understanding our
strate the ongoing relevance of evolutionary bi- cultural and genetic capacity for survival. The Pucuncho Basin in the southern Peruvian Andes
ology to contemporary environmental issues. contains the highest-altitude Pleistocene archaeological sites yet identified in the world,
about 900 meters above confidently dated contemporary sites. The Pucuncho workshop site
RE FE RENCES AND N OT ES
[4355 meters above sea level (masl)] includes two fishtail projectile points, which date to
1. W. L. Brown, E. O. Wilson, Syst. Zool. 5, 4964 (1956).
2. D. Schluter, The Ecology of Adaptive Radiation (Oxford Univ. about 12.8 to 11.5 thousand years ago (ka). Cuncaicha rock shelter (4480 masl) has a robust,
Press, Oxford, UK, 2000). well-preserved, and well-dated occupation sequence spanning the past 12.4 thousand years
3. T. Dayan, D. Simberloff, Ecol. Lett. 8, 875894 (2005). (ky), with 21 dates older than 11.5 ka. Our results demonstrate that despite cold temperatures
4. D. W. Pfennig, K. S. Pfennig, Evolution's Wedge (Univ. of
and low-oxygen conditions, hunter-gatherers colonized extreme high-altitude Andean environments
California Press, Berkeley, 2012).
5. Y. E. Stuart, J. B. Losos, Trends Ecol. Evol. 28, 402408 in the Terminal Pleistocene, within about 2 ky of the initial entry of humans to South America.

H
(2013).
6. P. R. Grant, B. R. Grant, Science 313, 224226 (2006). uman settlement of high-altitude moun- pecially prevalent in the treeless landscapes higher
7. J. G. Tyerman, M. Bertrand, C. C. Spencer, M. Doebeli, BMC tains and plateaus is among the most than 4000 meters above sea level (masl), with
Evol. Biol. 8, 34 (2008).
8. L. M. Bono, C. L. Gensel, D. W. Pfennig, C. L. Burch, Biol. Lett.
recent of our species biogeographic expan- little fuel for campfires, twice the sea-level caloric
9, 20120616 (2013). sions. Earths highest-altitude lands, located intake needed to maintain normal metabolic
9. M. L. Taper, in Bruchids and Legumes: Economics, Ecology, and in the Tibetan and Andean regions, pose function (2), and O2 partial pressure less than
Coevolution, K. Fujii, A. Gatehouse, C. D. Johnson, R. Mitchel, numerous physiological challenges, including 60% that at sea level (1). Current archaeological
T. Yoshida, Eds. (Kluwer, Dordrecht, Netherlands, 1990),
pp. 289301.
hypoxia (low-oxygen conditions), high solar ra- models (3) emphasize these challenges to explain
10. J. R. Edwards, S. P. Lailvaux, Biol. J. Linn. Soc. Lond. 110, diation, cold temperatures, and high energetic a lack of pre-Holocene [>11.5 thousand years ago
843851 (2013). costs of subsistence (1). These conditions are es- (ka)] (4) archaeological evidence above 4000 masl
11. J. B. Losos, Lizards in an Evolutionary Tree: Ecology and on the Tibetan (5) and Andean (6) Plateaus.
Adaptive Radiation of Anoles (Univ. of California Press,
Berkeley, 2009).
In the Andes, human biogeographic expansion
12. Information on materials and methods is available on Science 1
Department of Anthropology, South Stevens Hall, University to high-altitude lands likely stemmed from adja-
Online. of Maine, Orono, ME 04469-5773, USA. 2Department of Early cent areas in Peru (6), Chile (7), and Argentina (8)
13. T. W. Schoener, Ecol. Monogr. 45, 233258 (1975). Prehistory and Quaternary Ecology, Schlo Hohentbingen, (Fig. 1A). By ~13.5 to 12.1 ka or earlier, foragers
14. B. B. Collette, Bull. Mus. Comp. Zool. 125, 137162 (1961). Burgsteige 11, 72070 Tbingen, Germany. 3Accelerator Mass
15. L. R. Schettino et al., Breviora 520, 122 (2010). Spectrometry Laboratory, Department of Physics and School of
had settled the Pacific Coast (913) and the South-
16. J. R. Edwards, S. P. Lailvaux, Ethology 118, 494502 (2012). Anthropology, University of Arizona, Tucson, AZ 85721, USA. ern Cone (14), and by ~12.7 to 11.3 ka groups oc-
17. T. S. Campbell, thesis, University of Tennessee, Knoxville 4
University of Pennsylvania Museum, 3260 South Street, cupied caves at ~2600 masl in central Peru (15, 16)
(2000). Philadelphia, PA 19104, USA. 5Department of Anthropology and up to 3300 masl in the Atacama Desert of
18. D. Glossip, J. B. Losos, Herpetologica 53, 192199 (1997). and Archaeology, Earth Sciences Building, Room 806, 844
19. T. E. Macrini, D. J. Irschick, J. B. Losos, J. Herpetol. 37, 5258 Campus Place Northwest, Calgary, British Columbia, Canada.
northern Chile (17, 18). In northwest Argentina,
(2003). 6
Institute for Archaeological Sciences, University of Tbingen, multiple sites at 3400 to 3800 masl date to ~12.0 ka,
20. J. Elstrott, D. J. Irschick, Biol. J. Linn. Soc. Lond. 83, 389398 Rmelinstrasse 23, 72070 Tbingen, Germany. 7Senckenberg possibly as early as ~12.8 ka (8), although most pre-
(2004). Centre for Human Evolution and Paleoenvironment, University of Holocene occupations have only single, unrepli-
21. A. P. Hendry, M. T. Kinnison, Evolution 53, 16371653 Tbingen, Rmelinstrasse 23, 72070 Tbingen, Germany.
(1999). 8
Climate Change Institute, Bryand Global Sciences Center,
cated radiocarbon ages. Above 4000 masl, the
22. S. P. Carroll, C. Boyd, Evolution 46, 10521069 (1992). University of Maine, Orono, ME 04469, USA. 9Department of earliest known Andean sites (table S1) date from
23. D. N. Reznick, F. H. Shaw, F. H. Rodd, R. G. Shaw, Science 275, Anthropology, 354 Mansfield Road, University of Connecticut, the first millennium of the Holocene (19), with
19341937 (1997). Storrs, CT 06269-1176, USA. 10Department of Anthropology, widespread occupation after ~9 ka (68) and
24. S. Meiri, D. Simberloff, T. Dayan, J. Anim. Ecol. 80, 824834 University of Illinois at Chicago, Behavioral Sciences Building,
(2011). 1007 West Harrison Street, Chicago, IL 60607-7139, USA.
earliest year-round settlement after ~7.1 ka (20).
25. R. R. Tokarz, J. W. Beck Jr., Anim. Behav. 35, 722734 11
Arequipa, Peru. Whether genetic adaptations or environmen-
(1987). *Corresponding author. E-mail: kurt.rademaker@umit.maine.edu tal amelioration were necessary for high-altitude

466 24 OCTOBER 2014 VOL 346 ISSUE 6208 sciencemag.org SCIENCE


RE S EAR CH | R E P O R T S

human settlement remains poorly understood. sites yet identified anywhere in the world. These We discovered the Pucuncho and Cuncaicha
High-altitude regions are among the most remote sites extend the residence time of humans above archaeological sites (Fig. 1B and fig. S1) after map-
and least archaeologically studied lands on the 4000 masl by nearly a millennium, implying more ping of the Alca obsidian source (21), predictive
planet. Here, we report initial data from Terminal moderate late-glacial Andean environments and modeling (22), and systematic reconnaissance of
Pleistocene archaeological sites at nearly 4500 masl greater physiological capabilities for Pleistocene the volcanic plateau bounded by the Cotahuasi
in the Peruvian Andes, the highest Pleistocene humans than previously assumed. and Colca canyons of southern Peru (23). In the
heart of this plateau, the 132-km2 Pucuncho Ba-
sin contains grassland and wetland habitats and
thousands of domesticated camelids. The mod-
ern annual mean temperature at Pucuncho is 3C,
and prevailing easterly airflow maintains a semi-
arid climate, with 600 to 800 mm of annual pre-
cipitation primarily during the December to March
wet season (24).
The Pucuncho open-air workshop site (4355 masl)
on the basins west edge was situated to exploit
Alca-1 and Alca-5 obsidian pyroclasts eroding
from an alluvial fan. The workshop includes
debitage and 260 formal tools, including projec-
tile points and nondiagnostic bifaces, unifacial
scrapers, and other tools. Two fluted fishtail pro-
jectile point bases made of local fine-grained
andesite and Alca-5 obsidian (Fig. 2A) are diag-
nostic to ~12.8 to 11.5 ka (25); the Pucuncho fish-
tails are the highest fluted points yet discovered
in the Americas. Pucuncho likely provided the
Alca-1 obsidian in contexts dated ~13.4 to 10.2 ka
at Quebrada Jaguay, a Terminal Pleistocene fish-
ing settlement ~150 km south on the Pacific Coast
Fig. 1. Project area maps. (A) Andes showing Terminal Pleistocene and Early Holocene sites >2500 masl
(9) (Fig. 1B). Because no geologic mechanism can
and >4000 masl (also see table S1). (B) Study area.
transport Alca-1 obsidian to the coast, Quebrada
Jaguays inhabitants either visited the source or
obtained obsidian via exchange.
Seven km east of the Pucuncho site, the
Cuncaicha workshop site (4445 masl) occupies
an alluvial fan where Alca-5 and Alca-7 obsidian
pyroclasts crop out. This surface palimpsest con-
tains debitage and more than 500 projectile points
and nondiagnostic bifaces, scrapers, and other
unifacial tools, representing thousands of years
of episodic occupation. Above the alluvial fan is
Cuncaicha rock shelter (4480 masl), comprising
two north-facing alcoves formed by slab exfolia-
tion of the andesite bedrock. Both alcoves exhibit
sooted ceilings, rock art, and anthropogenic floor
sediments, indicating use as campsites. The rock
shelter is protected from westerly winds and of-
fers a commanding view of wetland and grass-
land habitats.
Our investigations at Cuncaicha rock shelter
sampled the minimum volume needed to docu-
ment the stratigraphy and establish a precise
absolute chronology of occupation. Ground-
penetrating radar revealed collapsed roof slabs
and depth of sediments to bedrock, allowing tar-
geted excavations. Sediments are ~1.2 m deep, both
within the shelter and outside the drip line over
~150 m2. We excavated 5.5 m2 (<4%) of these de-
posits in 2010 and 2012 (fig. S2).
We dated the Cuncaicha sequence using large-
mammal bone specimens in direct association
with abundant, unequivocal artifacts. Faunal and
other organic remains exhibit outstanding pres-
ervation in this cold, dry setting. Moreover, dating
Fig. 2. Terminal Pleistocene lithic artifacts. From (A) Pucuncho workshop (1 and 2) and (B and C) large bone specimens avoided the risk of sampling
Cuncaicha rock shelter (3 to 24) (also see table S4). (B) Projectile points (3 to 13) illustrated in strati- remains vertically translocated by rodent bio-
graphic order. (C) Selected tools and debitage (14 to 24) illustrating diverse forms and raw materials. turbation, a process shown to affect Andean rock

SCIENCE sciencemag.org 24 OCTOBER 2014 VOL 346 ISSUE 6208 467


R ES E A RC H | R E PO R TS

shelters elsewhere (15). Geoarchaeological analy- We recovered small, fragmented charred plant the quantity and diversity of early tool types; the
sis indicates only small-scale cryo- and bioturba- remains from sediments (26). Most abundant are emphasis on local lithic materials, animals, and
tion of deposits (26) (fig. S4). woody twigs, stems, and root wood, consistent combustible fuel; the presence of starchy roots
We obtained 35 accelerator mass spectrometry with burning small shrubs and Azorella compacta. and/or tubers; and the location in the heart of the
(AMS) ages at three laboratories using distinct Also present are charred fragments of parenchymous plateau suggest that Cuncaicha was a base camp.
pretreatment protocols on bone collagen (26) storage tissue (fig. S6). Their vitreous appearance The Pucuncho Basin constituted a high-altitude
(table S2). Dates on split samples at multiple indicates that these are starchy roots and/or oasis ideal for a specialized hunting (and later,
laboratories are statistically indistinguishable. tubers (30), likely gathered from lower elevations herding) adaptation. Vicua births coinciding with
The AMS-dated bone specimens are in correct and brought to the site for consumption (26). the end of the wet season and maintenance of
stratigraphic order, without reversals. Cuncaicha Cuncaicha is ~40 to 50 km from elevations permanent territories by vicua bands (29) would
rock shelter contains occupation components 2500 masl, so it is unlikely that the site was have permitted predictable scheduling of subsist-
corresponding with five distinct strata (fig. S3). merely a logistical station for the collection and ence activities and year-round plateau residence
Hiatuses correspond with clear stratigraphic sig- processing of lithic material, meat, and hides for (31). However, wet-season storms and the risk of
natures and are well constrained with AMS ages transport to low-elevation base camps. Together, hypothermia, as well as maintenance of extended
(Table 1 and table S2).
Feature 12-4 (an organic-rich pit containing ar-
tifacts) yielded the oldest reproduced ages, ~12.4 Table 1. Average AMS ages from Cuncaicha rock shelter. Component abbreviations: LMH, Late-Middle
to 11.8 ka (n = 2 AMS ages). Stratum 5 is com- Holocene; EH, Early Holocene; TP, Terminal Pleistocene. Two Late Holocene AMS ages were not averaged
posed largely of carbonates, likely derived from and are reported in table S2. Stratigraphic abbreviations: L, Level; F, Feature. CALIB 7.0.0 (39) reports the
anthropogenic burning of plants, including the test statistic T for a series of uncalibrated 14C ages from the same stratigraphic context (n) having a X2
local, resinous Azorella compacta (26) (fig. S5). distribution with n 1 degrees of freedom (df) under the null hypothesis of no difference with respect to a
Terminal Pleistocene-age deposits lack visible strat- threshold of a = 0.05 (40). Weighted means were calculated for statistically indistinguishable groups of
igraphic divisions. We grouped and averaged stratigraphically equivalent ages and calibrated using SHCal13 (41). 14C yr B.P., radiocarbon years before
AMS ages by excavation level (Table 1). The pos- the present. 68%/95% cal B.P., 68% and 95% probability age range in calendar years before the present.
itive age-depth relationship (Fig. 3) suggests epi-
sodic deposition ~12.4 to 11.4 ka (n = 19 AMS Component Context n df T 0.05 14
C yr B.P. 68% cal B.P. 95% cal B.P.
ages). The upper contact of stratum 5 is sharp LMH-II 3033 cm 3 2 1.2 6.0 4614 T 36 54465302 54665088
and distinct, likely formed during an occupation LMH-I 4045 cm 3 2 0.8 6.0 4867 T 37 56445586 56605484
hiatus ~11.4 to 9.5 ka. Stratum 4 corresponds to a EH 5062 cm 6 5 2.0 11.1 8420 T 34 94869432 95259324
brief, robust Early Holocene occupation ~9.5 to TP L.8-8b, 7779 cm 3 2 0.5 6.0 10,074 T 57 11,76411,406 11,96411,346
9.3 ka (n = 6 AMS ages), followed by a ~3.6 TP L.9, 7983 cm 3 2 0.2 6.0 10,073 T 49 11,76011,408 11,95811,357
thousand year (ky) mid-Holocene hiatus. Stra- TP L.11, 9098 cm 2 1 0.1 3.8 10,182 T 50 11,98211,769 12,06711,641
tum 3 includes distinct late-Middle Holocene oc-
TP L.12, 98108 cm 2 1 0.8 3.8 10,191 T 32 11,98011,822 12,03611,761
cupations dated ~5.7 to 5.5 ka (n = 3 AMS ages)
TP L.13, 108115 cm 7 6 7.8 12.6 10,228 T 18 11,95311,809 12,00411,759
and ~5.5 to 5.1 ka (n = 3 AMS ages). Strata 1 and
TP F.12-4 base 2 1 0.2 3.8 10,345 T 71 12,38611,953 12,41111,822
2 represent brief, episodic uses of the shelter
within the past ~2.2 ky (n = 2 AMS ages) (table S2).
Cuncaicha contains a rich assemblage of ceram- Fig. 3. Nevado
ics; chipped-stone tools, cores, and debitage; Coropuna 3He ages
faunal material; bone beads and quartz crystals; from C-II moraines
and fragments of red ochre (table S3). Ceramics, and average AMS
which locally date <4 ka (27), were found only in ages from Cuncaicha
Late Holocene strata 1 and 2. Temporal affilia- rock shelter Terminal
tions for the 153 projectile points found through- Pleistocene levels.
out the stratigraphy (23, 28) are consistent with Blue bar shows 95%
the AMS chronology. A complete lithic opera- range of C-II weighted
tional chain is present at Cuncaicha shelter and mean age.
the workshop below. Most lithic tools and debi-
tage at Cuncaicha are made from locally avail-
able Alca-1, -5, and -7 obsidian, andesite, and
jasper. Lithic tools indicate hunting and butchering
activities, consistent with the limited subsistence
options on the plateau.
The inhabitants of Cuncaicha hunted vicua
(Vicugna vicugna mensalis), guanaco (Lama
guanaco), and taruka (Hippocamelus antisensis)
(tables S5 and S6). Preliminary analysis of camelid
age profiles suggests predation at the end of the
rainy season when vicua are born (March and
April) and possibly during the dry season (May
to November) when vicua bands aggregate (29).
The even representation of mammal fore- and hind-
limb elements indicates dismembering of whole
carcasses at the shelter. First and second phalanges
are abundant, and the skinning of animals to the
toes attests to careful processing of all animal
foods, including meat and fat within the bone.

468 24 OCTOBER 2014 VOL 346 ISSUE 6208 sciencemag.org SCIENCE


RE S EAR CH | R E P O R T S

social networks and collection of edible plant re- 10. S. deFrance, N. Grayson, K. Wise, J. Field Archaeol. 34, 31. J. W. Rick, Prehistoric Hunters of the High Andes (Academic
sources, may have encouraged regular descents 227246 (2009). Press, New York, 1980).
11. D. Jackson, C. Mndez, R. Seguel, A. Maldonado, G. Vargas, 32. G. R. M. Bromley et al., Quat. Sci. Rev. 28, 25142526 (2009).
to lower elevations. Curr. Anthropol. 48, 725731 (2007). 33. P.-H. Blard et al., Quat. Sci. Rev. 30, 39733989 (2011).
Lithic tools and debitage of nonlocal fine- 12. T. D. Dillehay et al., Quat. Res. 77, 418423 (2012). 34. K. M. Weiss, A. W. Bigham, Evol. Anthropol. 16, 164171 (2007).
grained rocks, some with stream-polished cortex 13. T. D. Dillehay et al., Science 320, 784786 (2008). 35. A. R. Frisancho, Am. J. Hum. Biol. 25, 151168 (2013).
(Fig. 2C), suggest that Terminal Pleistocene and 14. J. Steele, G. Politis, J. Archaeol. Sci. 36, 419429 (2009). 36. A. W. Bigham et al., Am. J. Hum. Biol. 25, 190197 (2013).
15. T. F. Lynch, Guitarrero Cave: Early Man in the Andes (Academic 37. D. Zhou et al., Am. J. Hum. Genet. 93, 452462 (2013).
Early Holocene plateau residents ventured peri- Press, New York, 1980). 38. C. A. Eichstaedt et al., PLOS ONE 9, e93314 (2014).
odically to high-energy rivers below the plateau. 16. E. A. Jolie, T. F. Lynch, P. R. Geib, J. M. Adovasio, Curr. 39. M. Stuiver, P. J. Reimer, Radiocarbon 35, 215230 (1993).
Formal tools of Alca-4 obsidian at Cuncaicha orig- Anthropol. 52, 285296 (2011). 40. G. K. Ward, S. R. Wilson, Archaeometry 20, 1931 (1978).
inated in outcrops near the plateau edge ~22 km 17. M. Grosjean, L. Nez, I. Cartajena, J. Quat. Sci. 20, 643653 (2005). 41. A. G. Hogg et al., Radiocarbon 55, 18891903 (2013).
18. C. M. Santoro, L. Nez, Andean Past 1, 57109 (1987).
southwest (21) (Fig. 1B). Contemporary sites at 19. Singular Terminal Pleistocene dates from Peruvian sites AC KNOWLED GME NTS
Quebrada Jaguay on the Pacific Coast contain Pachamachay, Telarmachay, and PAn 12-58 are problematic Field work was supported by the Churchill Exploration Fund,
Alca-1, -4, and -5 obsidian tools and debitage (9, 23); (6) and generally rejected by the investigators; therefore, they National Geographic Society Waitt Foundation, Foundation for
the only source of these three obsidians is the are not included in this discussion. Exploration and Research on Cultural Origins, American
20. J. W. Rick, K. M. Moore, Boletn de Arqueologa 3, 263296 (1999). Philosophical Society, and Lambda Alpha. B. Robinson and B. Hall
Pucuncho Basin and surrounding plateau (21). The 21. K. Rademaker et al., Geology 41, 779782 (2013). funded pilot AMS dates. National Science Foundation Dissertation
oldest dates at Quebrada Jaguay and Cuncaicha 22. K. Rademaker, D. A. Reid, G. R. M. Bromley, in Least Cost Grant no. 1208748 funded the AMS chronology. Pucuncho Basin
overlap at two standard deviations. These sites Analysis of Social Landscapes: Archaeological Case Studies. pastoralists graciously allowed field work on their lands. A. Chu,
likely constitute end members in a coast-highland D. A. White, S. Surface-Evans, Eds. (University of Utah Press, R. Hintz, C. Mauricio, A. Ramos, A. Saenz, and E. Zuiga
Salt Lake City, UT, 2012), pp. 3244. provided invaluable logistical support. We thank field assistants
Paleoindian settlement system. 23. K. Rademaker, thesis, University of Maine, Orono (2012). W. Beckwith, K. Gardella, M. Koehler, T. Labanowski, O. McGlamery,
Pleistocene glaciers did not present a barrier to 24. U. Dornbusch, Erdkunde 52, 4154 (1998). E. Olson, and J. Wertheim and the staff of the Arizona AMS
human migration and settlement of the Pucuncho 25. L. J. Jackson, in Paleoindian Archaeology: A Hemispheric Laboratory. E. Cooper created Fig. 2. T. Koffman and P. Strand
Basin. Glacial-geologic records from adjacent Perspective. J. E. Morrow, C. Gnecco, Eds. (University of Florida calibrated C-II 3He ages for Fig. 3. Two anonymous reviewers
Press, Gainesville, 2006), pp. 105122. provided helpful comments on the manuscript. Additional data are
Nevado Coropuna (32) suggest that local glaciers 26. Materials and methods are available as supporting materials available in the supplementary materials. Artifact collections are
reached their late-Pleistocene maxima ~25 to 20 ka on Science Online. curated at the Ministry of Culture in Arequipa, Peru.
and even then did not encroach into the basin. 27. L. Perry et al., Nature 440, 7679 (2006).
After a relatively minor readvance ~13.4 ka (26) 28. C. J. Klink, M. S. Aldenderfer, in Advances in Titicaca Basin
Archaeology, vol. 1. C. Stanish, A. B. Cohen, M. S. Aldenderfer, SUPPLEMENTARY MATERIALS
(Fig. 3 and table S7), glaciers again receded. Eds. (Cotsen Institute of Archaeology at UCLA, Los Angeles,
Southward displacement of the intertropical con- www.sciencemag.org/content/346/6208/466/suppl/DC1
2005), pp. 2554.
Materials and Methods
vergence zone ~13.0 to 11.5 ka (33) probably re- 29. W. L. Franklin, in Mammalian Biology in South America,
Figs. S1 to S6
sulted in increased wet-season precipitation. The M. A. Mares, H. H. Genoways, Eds. (Volume 6, Special
Tables S1 to S7
Publication Series, Pymatuning Laboratory of Ecology,
arrival of humans to the Pucuncho Basin coincided University of Pittsburgh, Pittsburgh, 1981), pp. 457489.
References (4288)
with a period of warming climate and enhanced 30. J. G. Hather, Archaeological Parenchyma (Archetype 3 July 2014; accepted 16 September 2014
primary productivity in plateau habitats. Publications, London, 2000). 10.1126/science.1258260
Our data do not support previous hypotheses,
which suggested that climatic amelioration and
a lengthy period of human adaptation were nec-
essary for successful human colonization of the PLANT SCIENCE
high Andes. The Pucuncho Basin sites postdate
the oldest known South American lowland site,
Monte Verde (13), by only ~2 ky. Because early
settlement of high-altitude regions is understu-
Antheridiogen determines sex in
died, additional Terminal Pleistocene sites above
4000 masl likely await discovery. The Pucuncho
ferns via a spatiotemporally split
Basin sites suggest that Pleistocene humans lived
successfully at extreme high altitude, initiating
organismal selection (34), developmental function-
gibberellin synthesis pathway
al adaptations (35), and lasting biogeographic Junmu Tanaka,1* Kenji Yano,1* Koichiro Aya,1 Ko Hirano,1 Sayaka Takehara,1
expansion in the Andes. As new studies (3638) Eriko Koketsu,1 Reynante Lacsamana Ordonio,1 Seung-Hyun Park,2 Masatoshi Nakajima,2
identify potential genetic signatures of high- Miyako Ueguchi-Tanaka,1 Makoto Matsuoka1
altitude adaptation in modern Andean popula-
tions, comparative genomic, physiologic, and Some ferns possess the ability to control their sex ratio to maintain genetic variation in
archaeological research will be needed to under- their colony with the aid of antheridiogen pheromones, antheridium (male organ)inducing
stand when and how these adaptations evolved. compounds that are related to gibberellin. We determined that ferns have evolved an
antheridiogen-mediated communication system to produce males by modifying the
RE FE RENCES AND N OT ES
gibberellin biosynthetic pathway, which is split between two individuals of different
1. P. T. Baker, M. A. Little, Man in the Andes: Multidisciplinary
developmental stages in the colony. Antheridiogen acts as a bridge between them because
Study of High-Altitude Quechua (Dowden, Hutchinson, and
Ross, Stroudsberg, PA, 1976). it is more readily taken up by prothalli than bioactive gibberellin. The pathway initiates in
2. B. M. Marriot, S. J. Carlson-Newberry, Eds., Nutritional Needs in early-maturing prothalli (gametophytes) within a colony, which produce antheridiogens
Cold and High-Altitude Environments: Applications for Military and secrete them into the environment. After the secreted antheridiogen is absorbed by
Personnel in Field Operations (National Academy Press,
neighboring late-maturing prothalli, it is modified in to bioactive gibberellin to trigger male
Washington, DC, 1996).
3. M. S. Aldenderfer, World Archaeol. 38, 357370 (2006). organ formation.

G
4. All ages and ranges are two-sigma calibrated (2-s cal).
5. P. J. Brantingham et al., Geoarchaeol. 28, 413431 (2013). enetic diversity affords a competitive ad- within the population with the aid of antheridi-
6. K. Rademaker, G. R. M. Bromley, D. H. Sandweiss, Quat. Int. vantage to a particular species. Homo- ogens. Antheridiogens are pheromones released
301, 3445 (2013).
7. C. Mndez Melgar, Quat. Int. 301, 6073 (2013).
sporous ferns have evolved a mechanism to in the aqueous environment by early-maturing
8. L. Prates, G. Politis, J. Steele, Quat. Int. 301, 104122 (2013). favor cross-fertilization by controlling the fern prothalli in a colony, and they cause neigh-
9. D. H. Sandweiss et al., Science 281, 18301832 (1998). sex ratio among individuals or prothalli boring late-maturing prothalli in the colony to

SCIENCE sciencemag.org 24 OCTOBER 2014 VOL 346 ISSUE 6208 469


www.sciencemag.org/cgi/content/full/346/6208/466/DC1

Supplementary Materials for


Paleoindian settlement of the high-altitude Peruvian Andes
Kurt Rademaker,* Gregory Hodgins, Katherine Moore, Sonia Zarrillo,
Christopher Miller, Gordon R. M. Bromley, Peter Leach, David A. Reid,
Willy Ypez lvarez, Daniel H. Sandweiss

*Corresponding author. E-mail: kurt.rademaker@umit.maine.edu

Published 24 October 2014, Science 346, 466 (2014)


DOI: 10.1126/science.1258260

This PDF file includes:

Materials and Methods


Figs. S1 to S6
Tables S1 to S7
References
Materials and Methods
Arizona bone preparation protocol for accelerator mass spectrometry (AMS)

Approximately 500 mg of bone was crushed with a mortar and pestle and sieved to

recover a 0.5 mm to 1 mm diameter fraction. This material was loaded into a continuous flow

cell through which deionized water, 0.1 N HCl, 0.1 N NaOH, and 0.01 N HCl were pumped in a

programmed sequence that included both static incubation and flow. The recovered

demineralized collagen was then gelatinized at 70C for 20 hours in approximately 10 mls of pH

3.0 water, and then filtered through prewashed Whatman 0.45 micron glass microfiber Autovial

filters. This gelatin solution was then loaded into prewashed Sartorius Stedim Vivaspin 20,

30,000 dalton molecular weight cut-off ultrafilters and centrifuged at 900 x g for 15 minutes or

until the volume of the retained fraction was reduced to 0.5 to 1 ml. The retained fraction was

lyophilized to recover the ultrafiltered collagen for subsequent combustion, offline stable

isotope and C/N ratio analysis, followed by graphitization and AMS measurement by standard

methods.

Geoarchaeological investigations and soil micromorphology

In August 2012 Christopher Miller conducted a geoarchaeological study of Cuncaicha

and collected samples for micromorphological analysis. Micromorphology is the study of intact,

oriented blocks of sediment using microscopic techniques. This method allows us to not only

identify sedimentary components but also examine their structural relationships with each other.

We collected a total of nine block samples from Cuncaicha from all strata at the site,

focusing on the stratigraphic contacts. The samples were transported back to the

micromorphology laboratory at the University of Tbingen where they were dried and then

2
indurated with a mixture of unpromoted polyester resin, styrene, and Methyethylketone peroxide

under vacuum. Once the samples had achieved a gel-like consistency, they were heated

overnight at 60o C until completely hardened. The samples were then sliced with a rock saw and

made into thin sections measuring 70 mm x 90 mm and 30 m in thickness. The samples were

studied using a range of magnifications on a petrographic microscope (42, 43).

In the field, five distinct stratigraphic units were observed, including Stratum 1, which

consisted of a series of stacked combustion features; Stratum 2, a homogenous unit,

reddish/chestnut brown in color with isolated combustion features; Stratum 3, a dark, organic-

rich layer; Stratum 4, a homogenous, reddish-brown layer, and Stratum 5, which consisted of

greyish-colored sediments (see Fig. S3). The contacts between the units were clear but generally

diffuse, except for the contact between Strata 4 and 5, which was clear and sharp.

In thin section, the material from Strata 2, 3, and 4 exhibits a microgranular to granular

microstructure with chitonic c/f-related distribution; i.e., the grains are coated with isotropic

clays and finely comminuted organic material (Fig. S4). The coated grains are likely indicative

of cryoturbation resulting from repeated freeze and thaw of deposits (44). The microgranular

microstructure likely results from bioturbation by mesofauna; however, the clear macroscopic

preservation of strata suggests that bioturbation did not extensively homogenize the deposits.

Rather, movement of materials was likely limited to just a few centimeters (45). The centimeter-

scale movement of material is obvious at the contact between Stratum 4 and Stratum 5. In the

field this contact appeared clear and sharp. In thin section, this contact appeared diffuse over 2-3

centimetersa result of small-scale, mesofaunal bioturbation.

Stratum 5, which dates to the Pleistocene, is calcareous, unlike Strata 2, 3, and 4, which

are non-calcareous. Much of the calcite in Stratum 5 appears spherulitic, similar to the calcareous

3
ashes found in Stratum 1 (Fig. S5). Spherulitic calcite was also found in ashes produced by

experimentally burning Azorella compacta (llareta). Based on the comparison with ashes from

Stratum 1 and comparison with the morphology of experimentally produced ashes, it seems

likely that a significant portion of the calcareous deposits in Stratum 5 derive from the

anthropogenic burning of llareta plants.

Recovery of botanical remains and their identification

During the 2010 excavations, botanical remains were recovered from the 3 mm screen.

During the 2012 excavation, in order to test what may be missed in sediments that did not pass

through the 3 mm screen, the first four 10-liter buckets from each excavated context of a single

1.0 m x 0.5 m unit - Unit 7 - and for a few selected contexts from Units 3, 4 and 6 - were set

aside. We screened these ~40-liter samples using 3 mm screens as normal but also retained

everything remaining in the 3 mm screens. This typically resulted in a bagged sample of 1-2

liters per ~10-cm level, but in some cases these bagged samples were up to 10 liters.

These samples were then processed by water flotation to recover light and heavy

fractions. The dried heavy fractions were screened with standard USGS screens of >4mm, 2-4

mm, 0.5-2 mm, <0.5 mm and retained. The dried light fractions were weighed, sieved through a

2 mm geological screen and then manually sorted under low-power microscopy to recover

charred plant remains, lithic microdebitage, and faunal remains. The <2 mm splits were also

examined under microscopy, with these mostly consisting of fine sediments with very small

amounts of charred plant material too small to be identified. All of the >2 mm charred plant

remains were then examined further under higher magnification, separated into different classes,

and weighed.

4
Due to the different recovery methods employed during the 2010 and 2012 field season,

quantification and comparison across samples is not possible. The 2012 water flotation tests

show that preservation of charred botanical remains is excellent at the site.

Figure S6 shows examples of some of the charred parenchymous storage tissue

fragments recovered from Terminal Pleistocene contexts. Their tapered cylindrical shape, high

proportion of parenchyma tissue, and vitreous appearance are consistent with starchy roots

and/or tubers, i.e., underground storage organs (30, 46). More definitive identifications require a

broader assemblage of local and regional comparative taxa, as well as scanning electron

microscopy.

We believe that it is most likely that edible plant resources were obtained from lower

altitudes and brought to Cuncaicha because of the current flora and extreme conditions of the

Pucuncho Basin today, and the likelihood that similar or cooler conditions were present during

the Terminal Pleistocene occupation of the site. Starch-rich vegetative storage organs were and

are important dietary staples in the Andes today, especially at high altitudes where metabolic

requirements are extreme. As such, it is not surprising that Terminal Pleistocene people would

have collected plant resources at lower altitudes where they are most abundant and brought them

to the site for consumption. The fact that Terminal Pleistocene peoples in South America were

familiar with underground storage organs is evidenced by the recovery and identification of

Solanum maglia, a wild potato, from the Monte Verde site (47, 48).

Nevado Coropuna moraine records and paleoenvironmental interpretation

As sensitive indicators of climate, glaciers advance and retreat in response to small

changes primarily in temperature (49), leaving a physical record of these fluctuations in the form

5
of moraines. The moraine record from Coropuna is based on late-Pleistocene deposits on the

mountain's northern and western flanks and is resolved with cosmogenic 3He surface-exposure

dates (see 50 for a detailed description). The late-glacial moraine complex providing dates

reported here is located at ~5000 m elevation in Quebrada Santiago, a north-facing valley

draining into the Pucuncho Basin.

Of relevance to the present study is the late-glacial advance of Coropuna's glaciers that

interrupted gradual deglaciation during the last glacial-interglacial transition. This event, termed

the C-II advance (50), is represented in all valleys investigated on the mountain, as well as other

peaks in the Pucuncho region, and has been dated to ~12.9 ka. Since publication of that dataset,

however, further refinement of the scaling schemes used to calculate surface-exposure ages (51)

suggests that the scheme employed in that initial publication ('Lm' scaling; see 52) is less

accurate at these latitudes and altitudes than the 'St' scheme (52).

Consequently, we have recalculated the C-II moraine ages (50) using the 'St' scaling in

order to provide a more realistic age for the late-glacial advance on Nevado Coropuna (Table

S6). All other input data remain the same as in the original publication. Ten samples collected

from erratics located on the moraine crests in Quebrada Santiago (50) give recalculated ages

ranging from 12.0 to 13.9 ka, with a weighted mean age of 13.2 ka (Fig. 3). Plotted as an age-

probability curve, the dataset gives a peak (i.e., highest probability) age of 13.4 ka, in close

agreement with the weighted mean.

As in the original publication (50), we interpret the peak age as a close age for the C-II

moraines. Moreover, we reiterate that that this age represents the last time the glacier terminus

stood at this limit, rather than a date for the climate event that drove the glacial readvance. Thus,

we interpret the newly recalculated peak age of the C-II moraines as indicating retreat, and

6
therefore climatic amelioration, after ~13.4 ka. Based on this interpretation, we conclude that

initial human settlement of the Pucuncho Basin and Cuncaicha shelter occurred during a period

of resumed deglaciation.

In the context of the Cuncaicha Paleoindian occupation, our interpretation of the

Coropuna moraine record and other paleoclimate data from the southern Peruvian Andes is as

follows. First, as indicated by glacier reconstructions (32, 53), ice did not present a barrier to

human migration and settlement of the Pucuncho Basin. Indeed, local glaciers did not encroach

upon the plateau even at the height of the last ice age (~25-20 ka). Second, at the time of human

entry, glaciers in the Pucuncho region were retreating upslope in response to atmospheric

warming (50), a process that also would have reduced the areal extent of regional permafrost.

Third, between approximately 13 and 11.5 ka the inter-tropical convergence zone was displaced

to the south of its late-glacial position by cold stadial conditions in northern latitudes (54-58). As

a result, annual precipitation increased over much of the southern tropical Andes, as indicated by

elevated lake levels (e.g., 33, 58, 59). Therefore, it is highly likely that at the time of human

settlement, the Pucuncho Basin was experiencing relatively elevated wet-season precipitation,

resulting in a more vigorous hydrologic budget and greater primary productivity.

7
Fig. S1.
Photograph of Pucuncho Basin, facing west, showing Terminal Pleistocene sites Cuncaicha

shelter and workshop and Pucuncho workshop.

8
Fig. S2
Planview of Cuncaicha shelter with 2010 and 2012 excavations.

9
Fig. S3
Photograph of Cuncaicha rockshelter Unit 7 and 8 south profile showing stratigraphy and AMS

ages from Unit 7.

10
Fig. S4
Microphotograph of thin section (CUN-12-01) from Stratum 3, plane polarized light. The

microgranular microstructure and coated grains are clearly seen here. The microstructure and

coatings suggest that the deposits were subjected to limited bioturbation and cryoturbation.

11
Fig. S5
(A) Microphotograph of spherulitic ashes, produced by experimentally burning Azorella

compacta (llareta). The ashes were smeared onto a glass slide and photographed under cross-

polarized light (XPL). (B) Photomicrograph of thin section (CUN-12-09) from Stratum 5, XPL.

The spherulitic calcite strongly resembles that produced from the combustion of llareta. (C)

Microphotograph of thin section (CUN-12-01) from Stratum 1, which consists of stacked

combustion features. The ashy deposits from this stratum are composed of similar calcitic

spherulites as that found in Stratum 5, XPL.

12
Fig. S6
Examples of parenchymous storage tissue from Cuncaicha rockshelter Terminal Pleistocene

contexts. (A) and (C) from Unit 2, Level 8b, (B) from Unit 2, Level 10, (D) from Unit 7, Level

10a. (A), (B), and (D) are transverse sections; (C) is tangential/longitudinal. Scale bars = 5 mm.

13
Table S1.

Terminal Pleistocene and Early Holocene Andean sites >2500 masl.


14
Site Elevation C B.P. n= 68% cal BP 95% cal BP References
(masl) oldest age*
>4000 masl
Cuncaicha 4480 10,345 71 2 12,381-12,067 12,411-11,822 This work
Hakenasa 4100 9978 28 2 11,389-11,262 11,598-11,240 60
Hornillos-2 4020 9710 270 1 11,390-10,575 11,931-10,254 61
Panalauca 4150 9650 145 1 11,166-10,763 11,260-10,520 20
Quebrada Blanca 4080 9610 70 1 11,083-10,755 11,170-10,694 62
Las Cuevas 4540 9540 160 1 11,088-10,581 11,214-10,301 63
Pachamachay 4300 9010 285 1 10,409-9663 11,067-9425 31

3500-4000 masl
Pintoscayoc-1 3650 10,720 150 1 12,725-12,431 12,838-12,057 64
Inca Cueva-4 3730 10,620 140 1 12,698-12,178 12,723-12,041 65
Penas de las Trampas 3610 10,118 67 2 11,765-11,405 11,957-11,321 66
Tres Ventanas 3810 10,030 170 1 11,764-11,238 12,364-10,877 67
Quiqche-1 3600 9940 200 1 11,770-11,108 12,044-10,720 67
Huachichocana 3800 9340 120 1 10,650-10,295 11,063-10,220 68

3000-3500 masl
Cueva de Yavi 3430 10,450 55 1 12,419-12,106 12,541-12,021 69
Alero el Pescador 3300 10,310 130 1 12,390-11,765 12,543-11,405 70
Tuina-1 3160 10,220 117 2 12,050-11,414 12,390-11,325 71, 72
Tuina-5 3200 10,060 70 1 11,696-11,332 11,803-11,254 73
Quebrada Seca-3 3350 9790 50 1 11,232-11,147 11,254-10,877 74
Asana 3435 9683 98 2 11,181-10,791 11,226-10,710 75
Chulqui-1a 3280 9550 55 2 11,068-10,683 11,089-10,589 76
Lauricocha-2 3930 9525 260 1 11,173-10,439 11,616-9964 77
Jaywamachay 3400 9502 110 2 11,069-10,563 11,156-10,423 78

2500-3000 masl
Tulan-109 2950 10,700 94 2 12,701-12,455 12,737-12,418 73, 79
Salar de Punta Negra-1 2976 10,465 35 2 12,423-12,122 12,510-12,052 17
San Lorenzo-1 2950 10,281 88 2 12,362-11,754 12,407-11,509 80, 81
Guitarrero Cave 2580 10,262 43 2 11,994-11,824 12,057-11,753 15, 16
Salar de Punta Negra-6 2976 10,260 60 1 12,020-11,804 12,362-11,620 17
Agua de la Cueva 2906 10,248 58 2 11,998-11,773 12,251-11,618 82
Toquepala 2800 9490 140 1 11,070-10,513 11,171-10,299 83
14
* Oldest C ages are single, non-replicated ages (n=1) or weighted means of oldest statistically indistinguishable
ages (n=2).

14
Table S2. 14C AMS ages from Cuncaicha rockshelter bone samples.

Lab No. Context Depth (cm) 14


C B.P. 13C 95% cal BP Component
AA 101135 U.7, L.3 22-23.5 841 45 -19.7 788-664 LH II
AA 96338 U.2, L.3 31-41 2098 48 -19.0 2152-1900 LH I
AA 101133 U.6, L.4-5, F. 12-1 31-33 4584 59 -20.1 5447-4974 LMH II
AA 101134 U.6, L.4-5, F. 12-1 31-33 4683 73 -19.4 5584-5056 LMH II
AA 96340 U.2, L.3 31-41 4599 57 -19.1 5450-4978 LMH II
AA 96339 U.2, L.3 31-41 4826 58 -19.1 5641-5324 LMH I
AA 96335 U.2, L.4 41-52 4898 66 -19.6 5740-5332 LMH I
AA 101139 U.7, L.5 43-44 4890 66 -19.4 5732-5330 LMH I
AA 96337 U.1, L.4b 33-50 8363 82 -19.0 9492-9035 EH
AA 96336 U.2, L.4 41-52 8361 82 -18.9 9491-9034 EH
AA 96331 U.1-2, L.6 57-63 8404 82 -19.6 9523-9134 EH
AA 96329 U.1-2, L.6 57-63 8483 83 -19.9 9553-9146 EH
AA 101132 U.7, L.7, F.12-2 60-62 8454 84 -20.6 9543-9140 EH
AA 101131 U.7, L.7, F.12-2 60-62 8461 85 -19.8 9545-9141 EH
AA 96330 U.1-2, L.6 57-63 10,086 97 -19.3 11,953-11,253 TP
AA 101137 U.7, L.8, F.12-3 77-78 10,060 100 -20.1 11,945-11,238 TP
AA 96321 U.2, L.8b 69-79 10,034 97 -19.6 11,924-11,222 TP
AA 96322 U.2, L.8b 69-79 10,127 98 -19.9 11,991-11,282 TP
AA 94257 U.2, L.9 79-83 10,055 67 -19.7 11,793-11,250 TP
AA 96318 U.2, L.9 79-83 10,084 99 -19.7 11,954-11,250 TP
AA 96319 U.2, L.9 79-83 10,100 99 -19.2 11,962-11,266 TP
AA 96312 U.2, L.11 90-98 10,163 71 -18.9 12,009-11,393 TP
AA 94256 U.2, L.11 90-98 10,200 69 -20.1 12,039-11,405 TP
AA 96309 U.2, L.12 98-108 10,189 77 -19.4 12,035-11,399 TP
AA 94255 U.2, L.12 98-108 10,211 69 -20.5 12,051-11,406 TP
Beta 297423# U.2, L.13 108-115 10,050 50 -20.0 11,746-11,269 TP
AA 94254# U.2, L.13 108-115 10,321 73 -21.6 12,403-11,765 TP
AA 96306* U.2, L.13 108-115 10,132 71 -18.8 11,969-11,388 TP
PRI-12-029-04b* U.2, L.13 108-115 10,205 35 12,008-11,629 TP
AA 96307** U.2, L.13 108-115 10,306 72 -18.9 12,403-11,724 TP
PRI-12-029-05b** U.2, L.13 108-115 10,265 35 12,040-11,770 TP
AA 96308*** U.2, L.13 108-115 10,260 72 -19.2 12,384-11,502 TP
PRI-12-029-06b*** U.2, L.13 108-115 10,180 40 12,008-11,413 TP
AA 101138 U.7, L.10, F.12-4 105-108 10,310 100 -19.6 12,428-11,509 TP
AA 101130 U.7, L.10, F.12-4 102 10,380 100 -19.5 12,546-11,773 TP
Stratigraphic abbreviations: UUnit, LLevel, FFeature. Component abbreviations: LHLate Holocene, LMH-Late-Middle
Holocene, EHEarly Holocene, TPTerminal Pleistocene. All AMS measurements are corrected for 12C/13C isotopic
fractionation. AA samples are ultra-purified bone collagen measured at University of Arizona AMS Lab, Tucson, Arizona,
USA. Beta sample is bone collagen measured at Beta Analytic, Miami, Florida, USA. PRI samples are XAD ultra-purified
bone collagen prepared at PaleoResearch, Inc., Golden, Colorado, USA and measured at Keck Carbon Cycle AMS Facility,
University of California, Irvine, California, USA. Samples with # are splits prepared and analyzed at Beta Analytic and
Arizona. Samples with *, **, and *** are splits prepared at Arizona AMS and PaleoResearch, Inc. Calibrations were
performed using Calib 7.0.0 (39) and SHCal13 (41).

15
Table S3.

Excavated materials from Cuncaicha rockshelter.

Component Faunal Projectile Bifaces Scrapers Other Debitage Cores Beads, Ceramic
remains points n= n= tools wt (kg), n= n= Crystals wt (g), n=
(kg) n= n= n=
LH I-II 16.99 47 68 9 2 2.22, n=3655 0 4, 0 2347, n=454
LMH I-II 8.92 22 46 28 4 2.02, n=3041 0 4, 1 0, n=0
EH 23.69 54 45 78 1 2.47, n=3225 1 1, 1 0, n=0
TP 13.64 30 18 21 5 1.03, n=1570 2 8, 0 0, n=0
Total 63.24 153 177 136 12 7.74, n=11,491 3 17, 2 2347, n=454
Component abbreviations: LHLate Holocene, LMH-Late-Middle Holocene, EHEarly Holocene, TP Terminal Pleistocene.
Materials recovered from 5.5 m2 (4.5 m2 sampled Early Holocene deposits and 3.0 m2 sampled Terminal Pleistocene deposits).

16
Table S4.

Contexts and associated AMS ages for Cuncaicha rockshelter Terminal Pleistocene artifacts illustrated in

Figure 2.

Number Site Context Depth Tool type Raw AMS ages 95% ka
(cm) material (n=) range
A-1 Pucuncho surface 0 projectile point andesite
A-2 Pucuncho surface 0 projectile point obsidian
B-3 Cuncaicha U.2 L.8b 69-79 projectile point obsidian 2 12.0-11.3
B-4 Cuncaicha U.2 L.8b 69-79 projectile point obsidian 2 12.0-11.3
B-5 Cuncaicha U.7 L. 9 89 projectile point obsidian
B-6 Cuncaicha U.2 L.9 79-83 projectile point obsidian 3 12.0-11.4
B-7 Cuncaicha U.2 L.9 79-83 projectile point obsidian 3 12.0-11.4
B-8 Cuncaicha U.2 L.10 83-90 projectile point obsidian
B-9 Cuncaicha U.2 L.10 81 projectile point obsidian
B-10 Cuncaicha U.2 L.10 83-90 projectile point obsidian
B-11 Cuncaicha U.2 L.11 90-98 projectile point obsidian 2 12.1-11.6
B-12 Cuncaicha U.2 L.11 90-98 projectile point obsidian 2 12.1-11.6
B-13 Cuncaicha U.2 L.11 90-98 projectile point obsidian 2 12.1-11.6
C-14 Cuncaicha U.7 L.9 80-91 side scraper yellow chert
C-15 Cuncaicha U.7 L.10 103 heavy core tool red jasper
C-16 Cuncaicha U.8 L.10b 80-85 scraper/graver andesite
C-17 Cuncaicha U.2 L.8b 69-79 end scraper burned chert 2 12.0-11.3
C-18 Cuncaicha U.7 L.9 80-91 cortical flake green fine-grained
ign eous
C-19 Cuncaicha U.8 L.11 87-96 side scraper gray chert
C-20 Cuncaicha U.7 L.10a 91-114 end scraper white chalcedony
C-21 Cuncaicha U.6 L.8 66-87 flake fragment red-pink chert
C-22 Cuncaicha U.6 L.8 66-87 flake blue chalcedony
C-23 Cuncaicha U.2 L.10 87 end scraper obsidian
C-24 Cuncaicha U.5 L.11a 86-99 uniface andesite

17
Table S5.

Cuncaicha rockshelter taphonomic processes on faunal remains by component.

Component Observations Erosion Weathering Multiple Rodent Carnivore Burning- Burned-


(NISP) attritional damage damage calcined charred
agents
LH-I-II 37 16.2% 5.4% 2.7% present 14.3% 18.9%
LMH I-II 105 6.7% 18.1% 7.6% 1.0% present 4.8% 22.9%
EH 197 8.6% 7.1% 1.5% 2.5% present 9.1% 16.8%
TP 56 8.9% 7.1% 0.0% 3.6% 10.7% 16.1%
Component abbreviations: LHLate Holocene, LMH-Late-Middle Holocene, EHEarly Holocene, TPTerminal Pleistocene.
NISP: Number of identified specimens. Erosion is loss of material on surfaces and edges due to mechanical abrasion (84, p.
186); weathering is splitting and cracking of exposed bone surfaces (84, Behernsmeyers stage 2-4, Fig. 9.2); rodent damage
(84, Fig. 6.15); carnivore ravaging (84, Fig 6.20 and Fig. 6.21); multiple agents refers to weathering, followed by abrasion and
weathering, followed by burning. Burning as discussed in 85. The first five taphonomic agents probably occurred during the
time that the bone was exposed on an abandoned occupation surface, or when it was exposed to attritional agents after being
uncovered through rodent action or humans modifying an older surface. The burning processes occurred as a result of human
use of fire in heat features, cooking, or site maintenance. Calcined bone is produced at higher heats and more prolonged or
repeated heating. While a combination of non-cultural and cultural taphonomic changes is typical of all Cuncaicha rockshelter
faunal remains, the late-Middle to Late Holocene fauna show more evidence for weathering overall, and for exposure to more
diverse and intensive taphonomic alteration.

18
Table S6.

Large fauna from Cuncaicha rockshelter.

Component Elements Camelid Cervid Large Camelid NISP Camelid in Dental indicators
sampled (NISP) (NISP) Mammal % of identified vicua size
(NISP) (NISP) large mammal range
(n vicua)
vicua incisors
LH I-II 37 33 0 4 100.0% 50.0% (1)
present
LMH I-II 105 95 5 5 95.0% 61.1% (11)
non-vicua
EH 197 172 3 22 97.7% 73.3% (11) incisors present
(guanaco form)
TP 56 23 8 25 74.2% 50.0% (1)
Total 395 323 16 56
Component abbreviations: LHLate Holocene, LMH-Late-Middle Holocene, EHEarly Holocene, TPTerminal Pleistocene.
NISP: Number of identified specimens. Camelids and cervids were identified using the collections at the Zooarchaeology
Laboratory, University of Pennsylvania Museum. The camelid/cervid ratio is calculated based on all elements identifiable to
vertebrate family level. Camelid size range is determined for most abundant skeletal element (distal 1st phalanx) using
dimensions 4 (distal breadth) and 5 (distal depth), fore and hind limbs combined, based on (86, Tables 1 and 2). Dental
indicators for vicuna teeth following 87 and 88.

19
Table S7.

Cosmogenic 3He surface-exposure ages of ten glacial erratics located on Nevado Coropuna C-II moraines.

3
Sample ID He exposure age 1- 2- range

NC17 13,270 310 13,890-12,650

NC18 13,752 294 14,340-13,164

NC19 13,414 381 14,176-12,652

NC32 13,943 338 14,619-13,267

NC33 12,262 666 13,594-10,930

NC34 12,479 497 13,473-11,485

NC36 12,415 469 13,353-11,477

NC37 12,533 474 13,481-11,585

NC38 13,209 449 14,107-12,311

NC39 11,964 545 13,054-10,874

weighted mean 13,190 128 13,446-12,934

Ages recalculated from 50 using the 'St' scaling scheme (52).

20
References
1. P. T. Baker, M. A. Little, Man in the Andes: Multidisciplinary Study of High-Altitude Quechua
(Dowden, Hutchinson, and Ross, Stroudsberg, Pennsylvania, 1976).
2. B. M. Marriot, S. J. Carlson-Newberry, Eds., Nutritional Needs in Cold and High-altitude
Environments: Applications for Military Personnel in Field Operations (National
Academy Press, Washington, D.C., 1996).
3. M. S. Aldenderfer, Modelling plateau peoples: The early human use of the worlds high
plateau. World Archaeol. 38, 357370 (2006). doi:10.1080/00438240600813285
4. All ages and ranges are two-sigma calibrated (2- cal).
5. P. J. Brantingham, G. Xing, D. B. Madsen, D. Rhode, C. Perreault, J. van der Woerd, J. W.
Olsen, Late occupation of the high-elevation northern Tibetan Plateau based on
cosmogenic, luminescence, and radiocarbon ages. Geoarchaeol. 28, 413431 (2013).
doi:10.1002/gea.21448
6. K. Rademaker, G. R. M. Bromley, D. H. Sandweiss, Peru archaeological radiocarbon
database, 13,000-7000 14C B.P. Quat. Int. 301, 3445 (2013).
doi:10.1016/j.quaint.2012.08.2052
7. C. Mndez Melgar, Terminal Pleistocene/early Holocene 14C dates from archaeological sites
in Chile: Critical chronological issues for the initial peopling of the region. Quat. Int.
301, 6073 (2013). doi:10.1016/j.quaint.2012.04.003
8. L. Prates, G. Politis, J. Steele, Radiocarbon chronology of the early human occupation of
Argentina. Quat. Int. 301, 104122 (2013). doi:10.1016/j.quaint.2013.03.011
9. D. H. Sandweiss, H. McInnis, R. L. Burger, A. Cano, B. Ojeda, R. Paredes, M. C. Sandweiss,
M. D. Glascock, Quebrada Jaguay: Early South American maritime adaptations. Science
281, 18301832 (1998). Medline doi:10.1126/science.281.5384.1830
10. S. deFrance, N. Grayson, K. Wise, Documenting 12,000 years of coastal occupation on the
Osmore littoral, Peru. J. Field Archaeol. 34, 227246 (2009).
doi:10.1179/009346909791070853
11. D. Jackson, C. Mndez, R. Seguel, A. Maldonado, G. Vargas, Initial occupation of the
Pacific coast of Chile during late Pleistocene times. Curr. Anthropol. 48, 725731
(2007). doi:10.1086/520965
12. T. D. Dillehay, D. Bonavia, S. L. Goodbred Jr., M. Pino, V. Vsquez, T. R. Tham, A late
Pleistocene human presence at Huaca Prieta, Peru, and early Pacific Coastal adaptations.
Quat. Res. 77, 418423 (2012). doi:10.1016/j.yqres.2012.02.003
13. T. D. Dillehay, C. Ramrez, M. Pino, M. B. Collins, J. Rossen, J. D. Pino-Navarro, Monte
Verde: Seaweed, food, medicine, and the peopling of South America. Science 320, 784
786 (2008). Medline doi:10.1126/science.1156533
14. J. Steele, G. Politis, AMS 14C dating of early human occupation of southern South America.
J. Archaeol. Sci. 36, 419429 (2009). doi:10.1016/j.jas.2008.09.024
15. T. F. Lynch, Guitarrero Cave: Early Man in the Andes (Academic Press, New York, 1980).

21
16. E. A. Jolie, T. F. Lynch, P. R. Geib, J. M. Adovasio, Cordage, textiles, and the late
Pleistocene peopling of the Andes. Curr. Anthropol. 52, 285296 (2011).
doi:10.1086/659336
17. M. Grosjean, L. Nez, I. Cartajena, Palaeoindian occupation of the Atacama Desert,
northern Chile. J. Quat. Sci. 20, 643653 (2005). doi:10.1002/jqs.969
18. C. M. Santoro, L. Nez, Hunters of the dry puna and the salt puna in northern Chile.
Andean Past 1, 57109 (1987).
19. Singular Terminal Pleistocene dates from Peruvian sites Pachamachay, Telarmachay, and
PAn 12-58 are problematic (6) and generally rejected by the investigators; therefore, they
are not included in this discussion.
20. J. W. Rick, K. M. Moore, El Precermico de las punas de Junn: El punto de vista desde
Panaulauca. Boletn de Arqueologa 3, 263296 (1999).
21. K. Rademaker, M. D. Glascock, B. Kaiser, D. Gibson, D. R. Lux, M. G. Yates, Multi-
technique geochemical characterization of the Alca obsidian source, Peruvian Andes.
Geology 41, 779782 (2013). doi:10.1130/G34313.1
22. K. Rademaker, D. A. Reid, G. R. M. Bromley, 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. D. A. White, S.
Surface-Evans, Eds. (University of Utah Press, Salt Lake City, Utah, 2012), pp. 3244.
23. K. Rademaker, Early Human Settlement of the High-Altitude Pucuncho Basin, Peruvian
Andes (Ph.D. thesis, University of Maine, Orono, 2012).
24. U. Dornbusch, Current large-scale climatic conditions in southern Peru and their influence on
snowline altitudes. Erdkunde 52, 4154 (1998). doi:10.3112/erdkunde.1998.01.04
25. L. J. Jackson, Fluted and fishtail points from southern coastal Chile: New evidence
suggesting Clovis- and Folsom-related occupations in southernmost South America, in
Paleoindian Archaeology: A Hemispheric Perspective. J. E. Morrow, C. Gnecco, Eds.
(University of Florida Press, Gainesville, 2006), pp. 105122.
26. Materials and methods are available as supporting materials on Science Online.
27. L. Perry, D. H. Sandweiss, D. R. Piperno, K. Rademaker, M. A. Malpass, A. Umire, P. de la
Vera, Early maize agriculture and interzonal interaction in southern Peru. Nature 440,
7679 (2006). Medline doi:10.1038/nature04294
28. C. J. Klink, M. S. Aldenderfer, A projectile point chronology for the south-central Andean
highlands, in Advances in Titicaca Basin Archaeology, vol. 1. C. Stanish, A. B. Cohen,
M. S. Aldenderfer, Eds. (Cotsen Institute of Archaeology at UCLA, Los Angeles., 2005),
pp. 2554.
29. W. L. Franklin, Biology, ecology, and relationship to man of the South American camelids,
in Mammalian Biology in South America, M. A. Mares, H. H. Genoways, Eds. (Volume
6, Special Publication Series, Pymatuning Laboratory of Ecology, University of
Pittsburgh, 1981), pp. 457489.
30. J. G. Hather, Archaeological Parenchyma (Archetype Publications, London, 2000).

22
31. J. W. Rick, Prehistoric Hunters of the High Andes (Academic Press, New York, 1980).
32. G. R. M. Bromley, J. M. Schaefer, G. Winckler, B. L. Hall, C. E. Todd, K. M. Rademaker,
Relative timing of last glacial maximum and late-glacial events in the central tropical
Andes. Quat. Sci. Rev. 28, 25142526 (2009). doi:10.1016/j.quascirev.2009.05.012
33. P.-H. Blard, F. Sylvestre, A. K. Tripati, C. Claude, C. Causse, A. Coudrain, T. Condom, J.-L.
Seidel, F. Vimeux, C. Moreau, J.-P. Dumoulin, J. Lav, Lake highstands on the altiplano
(tropical Andes) contemporaneous with Heinrich 1 and the Younger Dryas: New insights
from 14C, U-Th dating and 18O of carbonates. Quat. Sci. Rev. 30, 39733989 (2011).
doi:10.1016/j.quascirev.2011.11.001
34. K. M. Weiss, A. W. Bigham, So mortal and so strange a pang: A tribute to Paul T. Baker.
Evol. Anthropol. 16, 164171 (2007). doi:10.1002/evan.20140
35. A. R. Frisancho, Developmental functional adaptation to high altitude: Review. Am. J. Hum.
Biol. 25, 151168 (2013). Medline doi:10.1002/ajhb.22367
36. A. W. Bigham, M. J. Wilson, C. G. Julian, M. Kiyamu, E. Vargas, F. Leon-Velarde, M.
Rivera-Chira, C. Rodriquez, V. A. Browne, E. Parra, T. D. Brutsaert, L. G. Moore, M. D.
Shriver, Andean and Tibetan patterns of adaptation to high altitude. Am. J. Hum. Biol. 25,
190197 (2013). Medline doi:10.1002/ajhb.22358
37. D. Zhou, N. Udpa, R. Ronen, T. Stobdan, J. Liang, O. Appenzeller, H. W. Zhao, Y. Yin, Y.
Du, L. Guo, R. Cao, Y. Wang, X. Jin, C. Huang, W. Jia, D. Cao, G. Guo, J. L. Gamboa,
F. Villafuerte, D. Callacondo, J. Xue, S. Liu, K. A. Frazer, Y. Li, V. Bafna, G. G.
Haddad, Whole-genome sequencing uncovers the genetic basis of chronic mountain
sickness in Andean highlanders. Am. J. Hum. Genet. 93, 452462 (2013). Medline
doi:10.1016/j.ajhg.2013.07.011
38. C. A. Eichstaedt, T. Anto, L. Pagani, A. Cardona, T. Kivisild, M. Mormina, The Andean
adaptive toolkit to counteract high altitude maladaptation: Genome-wide and phenotypic
analysis of the Collas. PLOS ONE 9, e93314 (2014). Medline
doi:10.1371/journal.pone.0093314
39. M. Stuiver, P. J. Reimer, Extended 14C data base and revised CALIB 3.0 14C age calibration
program. Radiocarbon 35, 215230 (1993).
40. G. K. Ward, S. R. Wilson, Procedures for comparing and combining radiocarbon age
determinations: A critique. Archaeometry 20, 1931 (1978). doi:10.1111/j.1475-
4754.1978.tb00208.x
41. A. G. Hogg, Q. Hua, P. G. Blackwell, M. Niu, C. E. Buck, T. P. Guilderson, T. J. Heaton, J.
G. Palmer, P. J. Reimer, R. W. Reimer, C. S. M. Turney, S. R. H. Zimmerman, SHCal13
southern hemisphere calibration, 050,000 years cal BP. Radiocarbon 55, 18891903
(2013). doi:10.2458/azu_js_rc.55.16783
42. M. A. Courty, P. Goldberg, P. Macphail, Soils and Micromorphology in Archaeology
(Cambridge Univ. Press, Cambridge, UK, 1989).
43. G. Stoops, Guidelines for Analysis and Description of Soil and Regolith Thin Sections (Soil
Science Society of America, Madison, Wisconsin, 2003).

23
44. P. Goldberg, S. Schiegl, K. Meligne, C. Dayton, N. J. Conard, Micromorphology and site
formation at Hohle Fels Cave, Swabian Jura, Germany. Eiszeitalter und Gegenwart 53,
125 (2003).
45. S. Mentzer, Macro- and Micro-scale Geoarchaeology of azl Caves I and II, Hatay,
Turkey (Ph.D. Dissertation, University of Arizona, Tucson, 2011).
46. D. M. Pearsall, Paleoethnobotany: A Handbook of Procedures (Academic Press, San Diego,
2000).
47. D. Ugent, The tuberous plant remains of Monte Verde, in Monte Verde: A Late Pleistocene
Settlement in Chile: The Archaeological Context and Interpretations, vol. 2., T. D.
Dillehay (Smithsonian Institution Press, Washington, D.C., 1997), pp. 903910.
48. D. Ugent, T. Dillehay, C. Ramirez, Potato remains from a late-Pleistocene settlement in
south-central Chile. Econ. Bot. 41, 1727 (1987). doi:10.1007/BF02859340
49. J. Oerlemans, Glaciers and Climate Change (A. A. Balkema Publishers, Rotterdam, 2001).
50. G. R. M. Bromley, B. L. Hall, J. M. Schaefer, G. Winckler, C. E. Todd, K. M. Rademaker,
Glacier fluctuations in the southern Peruvian Andes during the late-glacial period,
constrained with cosmogenic 3He. J. Quat. Sci. 26, 3743 (2011). doi:10.1002/jqs.1424
51. M. A. Kelly et al., A locally calibrated, late glacial 10Be production rate from a low-latitude,
high-altitude site in the Peruvian Andes, Quat. Geochron.; doi:
10.1016/j.quageo.2013.10.007 (2013).
52. G. Balco, J. O. Stone, N. A. Lifton, T. J. Dunai, A complete and easily accessible means of
calculating surface exposure ages of erosion rates from 10Be and 26Al measurements.
Quat. Geochronol. 3, 174195 (2008). doi:10.1016/j.quageo.2007.12.001
53. G. R. M. Bromley, B. L. Hall, K. M. Rademaker, C. E. Todd, A. E. Racovteanu, Late
Pleistocene snowline fluctuations at Nevado Coropuna (15S), southern Peruvian Andes.
J. Quat. Sci. 26, 305317 (2011). doi:10.1002/jqs.1455
54. L. C. Peterson, G. H. Haug, K. A. Hughen, U. Rhl, Rapid changes in the hydrologic cycle of
the tropical Atlantic during the last glacial. Science 290, 19471951 (2000). Medline
doi:10.1126/science.290.5498.1947
55. Y. J. Wang, H. Cheng, R. L. Edwards, Z. S. An, J. Y. Wu, C. C. Shen, J. A. Dorale, A high-
resolution absolute-dated late Pleistocene Monsoon record from Hulu Cave, China.
Science 294, 23452348 (2001). Medline doi:10.1126/science.1064618
56. X. Wang, A. S. Auler, R. L. Edwards, H. Cheng, P. S. Cristalli, P. L. Smart, D. A. Richards,
C. C. Shen, Wet periods in northeastern Brazil over the past 210 kyr linked to distant
climate anomalies. Nature 432, 740743 (2004). Medline doi:10.1038/nature03067
57. J. C. H. Chiang, C. M. Bitz, Influence of high latitude ice cover on the marine Intertropical
Convergence Zone. Clim. Dyn. 25, 477496 (2005). doi:10.1007/s00382-005-0040-5
58. C. Placzek, J. Quade, P. J. Patchett, Geochronology and stratigraphy of late Pleistocene lake
cycles on the southern Bolivian Altiplano: Implications for causes of tropical climate
change. Geol. Soc. Am. Bull. 118, 515532 (2006). doi:10.1130/B25770.1

24
59. P. A. Baker, C. A. Rigsby, G. O. Seltzer, S. C. Fritz, T. K. Lowenstein, N. P. Bacher, C.
Veliz, Tropical climate changes at millennial and orbital timescales on the Bolivian
Altiplano. Nature 409, 698701 (2001). Medline doi:10.1038/35055524
60. D. Osorio, D. Jackson, P. C. Ugalde, C. Latorre, R. De Pol-Holz, C. M. Santoro, Hakenasa
Cave and its relevance for the peopling of the southern Andean altiplano. Antiquity 85,
11941208 (2011).
61. H. Yacobaccio, M. P. Cat, P. Sol, M. Alonso, Estudio arqueolgico y fsicoqumico de
pinturas rupestres en Hornillos 2 (Puna de Jujuy). Estud. Atacameos 36, 528 (2008).
62. M. Grosjean, C. M. Santoro, L. G. Thompson, L. Nunez, V. Standen, Mid-Holocene climate
and culture change in the south-central Andes, in Climate Change and Cultural
Dynamics: A Global Perspective on Mid-Holocene Transitions, D. G. Anderson, K. A.
Maasch, D. H. Sandweiss (Academic Press, Burlington, Massachusetts, 2007), pp. 51
116.
63. C. Santoro, J. Chacama, Secuencia de asentamientos precermicos del extremo norte de
Chile. Estud. Atacameos 7, 85103 (1984).
64. M. I. Hernndez Llosas, Quebradas altas de Humahuaca a travs del tiempo: El caso
Pintoscayoc. Estudios Sociales del NOA 2, 167224 (2000).
65. C. Aschero, M. Podest, El arte rupestre en asentamientos precermicos de la Puna
Argentina. Runa 16, 2957 (1986).
66. J. G. Martnez, M. Mondini, E. Pintar, M. C. Reigadas, Cazadores-recolectores tempranos de
la Puna meridional Argentina: Avances en su estudio en Antofagasta de la Sierra
(Pleistoceno final-Holoceno temprano/medio), in Proceedings of the 17th Congreso
Nacional de Arqueologa Argentina, vol. III (Facultad de Universidad Nacional de Cuyo,
Mendoza, 2010), pp. 16911696.
67. F. Engel, S. Genoves, On early man in the Americas. Curr. Anthropol. 10, 225 (1969).
doi:10.1086/201078
68. C. Aschero, El sitio IC-C4: Un asentamiento precermico en la Quebrada de Inca Cueva
(Jujuy, Argentina). Estud. Atacameos 7, 6272 (1984).
69. J. A. Kulemeyer, L. R. Laguna, La Cueva de Yavi: Cazadores-recolectores del borde oriental
de la Puna de Jujuy (Argentina) entre los 12,500 y 8,000 aos 14C BP. Ciencia y
Tecnologa l, 37 (1996).
70. P. DeSouza, Cazadores recolectores del Arcaico Temprano y Medio en la cuenca superior del
rio Loa: Sitios, conjuntos lticos y sistemas de asentamiento. Estud. Atacameos 27, 743
(2004).
71. J. L. Betancourt, C. Latorre, J. A. Rech, J. Quade, K. A. Rylander, A 22,000-year record of
monsoonal precipitation from northern Chiles Atacama Desert. Science 289, 15421546
(2000). Medline doi:10.1126/science.289.5484.1542
72. C. Latorre, J. L. Betancourt, K. A. Rylander, J. Quade, Vegetation invasions into absolute
desert: A 45 000 yr rodent midden record from CalamaSalar de Atacama basins,
northern Chile (lat. 2224 S). Geol. Soc. Am. Bull. 114, 349366 (2002).
doi:10.1130/0016-7606(2002)114<0349:VIIADA>2.0.CO;2

25
73. L. Nez, M. Grosjean, I. Cartajena, Human occupations and climate change in the Puna de
Atacama, Chile. Science 298, 821824 (2002). Medline doi:10.1126/science.1076449
74. S. Hocsman, Cazadores-recolectores complejos en la Puna meridional Argentina?
Entrelazando evidencias del registro arqueolgico de la microrregin de Antofagasta de
la Sierra, Catamarca. Relaciones de la Sociedad Argentina de Antropologa 27, 193214
(2002).
75. M. S. Aldenderfer, Montane Foragers: Asana and the South-Central Andean Archaic
(University of Iowa Press, Iowa City, 1998).
76. C. Sinclair, Dos fechas radiocarbonicas del alero Chulqui, rio Toconce: Noticia y comentario.
Chungara 14, 7179 (1985).
77. A. Cardich, Lauricocha: Fundamentos para una prehistoria de los Andes centrales. Acta
Prehistrica 8/10, 3171 (1960).
78. R. S. MacNeish, A. G. Cook, L. G. Lumbreras, R. K. Vierra, A. Nelken-Turner, Prehistory of
the Ayacucho Basin, Peru, vol. II, Excavations and Chronology (University of Michigan
Press, Ann Arbor, 1981).
79. L. Nez, M. Grosjean, I. Cartajena, I., Ocupaciones Humanas y Paleoambientes en la Puna
de Atacama (Instituto de Investigaciones Arqueolgicas y Museo, Universidad Catlica
del Norte Taraxacum, San Pedro de Atacama, 2005).
80. J.-C. Spahny, Recherches archologiques a lembouchure du rio Loa (Cte du Pacifique-
Chili). J. Soc. Am. 56, 179251 (1967). doi:10.3406/jsa.1967.2276
81. L. Nez, Paleoindian and Archaic cultural periods in the arid and semiarid regions of
northern Chile, in Advances in World Archaeology, vol. II., F. Wendorf, A. C. Close
(Academic Press, New York, 1983), pp. 161201.
82. A. Garca, M. Zrate, M. Pez, The Pleistocene Holocene transition and the human
occupation in the Central Andes of Argentina: Agua de la Cueva locality. Quat. Int. 53-
54, 4352 (1999). doi:10.1016/S1040-6182(98)00006-8
83. R. Ravines, Investigaciones arqueolgicas en el Per: 1965-1966. Rev. Mus. Nac. 34, 247
254 (1966).
84. R. L. Lyman, Vertebrate Taphonomy (Cambridge Univ. Press, Cambridge, 1996).
85. K. M. Moore, M. Bruno, J. Capriles, C. Hastorf, Integrated contextual approaches to
understanding past activities using plant and animal remains from Kala Uyuni (Lake
Titicaca, Bolivia), in Integrating Zooarchaeology and Paleoethnobotany: A
Consideration of Issues, Methods, and Cases, A. M. VanDerwarker, T. M. Peres
(Springer Verlag, New York, 2010).
86. A. D. Izeta, C. Otaola, A. Gasco, Osteometra de falanges proximales de camlidos
sudamericanos modernos. Variabilidad, estndares mtricos y su importancia como
conjunto comparativo para la interpretacin de restos hallados en contextos
arqueolgicos. Rev. Mus. Antropol. 2, 169180 (2009).

26
87. J. C. Wheeler, De la Chasse dElevage, in Telarmachay: Chasseurs et Pasteurs
Prhistoriques des Andes, vol. I, D. Lavalle, M. Julien, J. Wheeler, C. Karlin (Editions
Recherche sur les Civilisations, Synthse 20, Paris, 1985), pp. 6179.
88. G. L. Mengoni Goalons, H. D. Yacobaccio, The domestication of South American
camelids: A view from the south-central Andes, in Documenting Domestication: New
Genetic and Archaeological Paradigms, M. E. Zeder, D. G. Bradley, E. Emshwiller, B.
D. Smith (University of California Press, Berkeley, 2006), pp. 228244.

27

Das könnte Ihnen auch gefallen