Anda di halaman 1dari 133

Quaternary International 363 (2015) 1e3

Contents lists available at ScienceDirect

Quaternary International
journal homepage: www.elsevier.com/locate/quaint

Guest Editorial

Multidisciplinary studies on the humaneenvironment interaction during


the initial peopling of the Americas
Climatic and environmental impact on humans and their role in
animal and vegetation communities during the Pleistocene/Holocene transition (ca. 13,000e8000 14C BP) is a major topic for the
archaeology of the initial human peopling of the Americas. This
time period is crucial to understand the range of adaptive strategies
displayed by early hunter-gatherers to changing ecosystems and
climatic conditions during their expansion throughout the continent. The detailed chronoclimatic framework reveals high climatic
uctuation (in temperature, humidity, dryness), whereas the
archaeological record suggests signicant cultural diversity, exemplied by the appearance of various techno-complexes dispersed
through space and time, diversication of subsistence resources
with special emphasis on plants, and broad spectrum economies.
Thus, the complex humaneenvironment interaction has been a
recurrent issue in the research agenda. However, given the fragmentary archaeological and paleoecological record, such interactions are difcult to address at a continental scale.
Notwithstanding that climaticehumaneenvironment interaction is by denition dynamic, past interpretations have both characterized early hunteregatherers as mere resource consumers
and overemphasized, from a deterministic perspective, the role of
climatic and environmental events over human societies. Nevertheless, new ideas and the supporting evidence suggest that human
activities and environmental change should be viewed together as
a coevolutionary and adaptive process. This change in understanding of this interaction was possible because of the contribution of
distinct disciplines (archaeology, paleoecology, taphonomy, paleontology, bioarchaeology, etc) and evidence (lithic technology, paleoclimatic and paleoenvironmental proxies, radiometric
databases, faunal and human skeletal remains, among others)
investigated from different spatial scales (from micro to macro
scales).
Here, we present an invited volume in Quaternary International
integrated by a series of papers that deal with the humaneenvironment interaction during the initial peopling phase from a continental perspective from Mexico to Patagonia, including the Pacic coast
of South America, the Andean highlands, the tropical lowlands, and
the wide grasslands of the south of the continent. The evidence
studied and the methodological and theoretical approaches were
diverse. This volume is the result of the session chaired by Miguel
Delgado and Francisco Aceituno entitled, Multidisciplinary studies
on the humaneenvironment interaction during the initial phase of
the American peopling held during the VI International Symposium Early Man in America: Models and Contributions from Tropical Territories presented in Pereira, Colombia, 19e24 November,
2012. The main aims of the session were to offer a place to present
recent results favoring multidisciplinary approaches on the
http://dx.doi.org/10.1016/j.quaint.2015.01.040
1040-6182/ 2015 Elsevier Ltd and INQUA. All rights reserved.

interaction between humans, the environment (climate and landscape) and other animal and vegetation species during the Pleistocene/Holocene transition and to discuss new theoretical and
methodological strategies to address interaction from the archaeological record in relation with other evidence.
The ten papers included in this volume are representative of the
current diversity of research topics and methodological/theoretical
perspectives on the archaeological study of the humaneenvironment interaction during the Pleistocene/Holocene transition in
the Americas. Together, the papers demonstrate at different levels
and spatiotemporal scales the complex dynamics between humans
and the environment, integrated by several biotic and abiotic
agents that potentially promoted their biocultural evolution. The
case studies cover a wide spectrum of contexts, evidence, and approaches focused on hunteregatherers, spanning from the late
Pleistocene to the early/middle Holocene. Interestingly, two papers
included addressed aspects of total relevance such as plant domestication using experimental approaches on and the development of
theoretical perspectives to study early American hunter-gatherers.
ntara, G. DomThe work of S. Gonzalez, D. Huddart, I. Israde-Alca
squez, J. Bischoff and N. Felstead presents new data
nguez-Va
regarding stratigraphy, dating, and tephrochronology of some
important late Pleistocene/early Holocene sites in the Mexico Basin.
Their results indicate the likely inuence of volcanic processes and
a meteorite airburst event on human, animal and vegetation communities during the Pleistocene/Holocene transition. This paper
enhanced the chronological resolution of the studied sites, an
important step to discuss the initial human peopling of the Mexico
basin, and provides detailed evidence on the multiple events that
likely inuenced the early hunter-gatherer way of life and dispersals throughout the region. Importantly, this work stressed the
relevance of human skeletal remains recovered, which have some
of the earliest radiocarbon dates obtained in the Americas.
The four following papers concern the environmentally and
culturally diverse northwestern South American region, including
much of the current Colombian territory. They all emphasize adaptive strategies mostly based on plant resources, the lithic technology, and the chronology of the early archaeological record.
Moreover, two of these papers present both new radiometric measurements of some archaeological contexts and critical evaluations
of the regional chronological database within an archaeological and
paleoenvironmental framework.
The article of F. Aceituno and N. Loaiza focuses on the adaptive
strategies of human groups that settled Northwest South America
(Colombia) during the Pleistocene/Holocene transition. The authors
present a synthesis that suggests the key role of plant resources in
the human settling of neotropical forests. The evidence analyzed,

Guest Editorial / Quaternary International 363 (2015) 1e3

basically lithic technology and the archaeobotanical record, support


previous ideas on the impact of early hunter-gatherer societies on
the environment through forest clearing, burning, and cultural selection of key vegetation resources. In addition, this paper contributes to the discussion of early plant domestication and cultivation
in tropical America.
G. Santos, C. Monsalve, and L. Correa present data from the
Medelln-River Basin in the Cordillera Central (Colombia) which
suggest that the human disturbance of the tropical forest increased
the carrying capacity. Furthermore, there is evidence that reveals
the importance of the small-scale cultivation of plants since the
end of the early Holocene. This work highlights the importance of
anthropogenic landscapes used by hunteregatherers during the
peopling of Northwest South America.
 pez, M. Cano, L. Herrera, C.
R. Dickau, F. Aceituno, N. Loaiza, C. Lo
Restrepo, and A. Ranere reported new late Pleistocene/middle Holocene sites (between 10,600 and 3600 14C BP) lying in the Middle
Cauca, in the Cordillera Central (Colombia). In this article, the authors described new sites and presented a solid radiocarbon chronology of 26 sites in order to obtain a better chronological
resolution on the peopling of Northwestern South America. This
work also includes the description of lithic technology and the
archaeobotanical evidence for the early plant use in this neotropic
region and contributes to the clarication of cultural periodization
on the basis of archaeological materials.
M. Delgado, F. Aceituno, and G. Barrientos presented the
most updated regional analysis of the radiometric database
corresponding to the Pleistocene/Holocene transition (ca.
13,000e8000 14C BP) in order to establish a reliable chronological
framework, critical to inferences on the timing, pattern, process
and tempo of early exploration and colonization of the study
area. The authors using uncalibrated and calibrated dates explored
some spatial and temporal trends illustrating the regions with the
earliest and latest occupations within the chronological range
considered and the uctuation of the archaeological signal before
and after 11,000 14C BP. Using paleoclimatic proxies and published
paleoenvironmental reconstructions considering pre-11,000 14C
ages, the authors suggested that the exploration phase occurred
during a relatively warm and humid period, but it was during
the El Abra Stadial, contemporary with the Younger Dryas Chronozone (12,900e11,600 cal BP) when the colonizing population rst
reached an indisputably clear archaeological visibility represented
by a diversity of contexts deposited at different environmental settings. These four articles represent a major contribution for the
study of the early human occupations in the northwestern South
American region that must be inserted in broad discussions at a
continental level.
The paper of D. Piperno, I. Holst, K. Winter, and O. McMillan presents the results of an excellent experiment on the domestication of
teosinte (Zea mays ssp. parviglumis), specically on phenotypic diversity and productivity (biomass, seed yield). The experimental
data suggest that the environmental conditions of the terminal
Pleistocene and early Holocene were a crucial factor to explain
plant cultivation origins, such as teosinte. For instance, the lower
level of Pleistocene CO2, temperature and precipitation uctuations
were signicant limiting factors on plant growth compared with
the following Holocene epoch. In contrast, the increase of Holocene
CO2 favored teosinte productivity and the appearance of new
phenotypic traits, fact that was taken advantage of by the human
populations. Probably from this moment, the foragers would focus
their attention on more productive plants that later were cultivated. This work highlights the potential of new methodological
tools derived from actualistic approaches to address important issues such as New World agriculture emergence following the end
of the Pleistocene.

The following three papers are framed on southern South American environmental scenarios from the dry Puna to Pampa and
Patagonia grasslands in Argentina. They focus on the initial human
colonization, paleoenvironmental scenarios, and lithic technology.
Some of these papers addressed interesting issues such as megafaunal extinction, water availability in arid and semiarid environments, chemical analysis of consumed resources, and lithic
taphonomy.
The work of R. Hoguin and B. Oxman describes the relation between climatic/environmental uctuations, resource availability,
and the lithic technological strategies followed by early hunteregatherers during the initial peopling of the dry Puna in northwestern Argentina. In this case-study, the authors use pollen
analysis to suggest that the occurrence of humid conditions,
although not synchronously, during the early Holocene favored
the extension of wetlands and the expansion of Andean grasslands,
allowing the increase of regional carrying capacity. The huntergatherer response was to increase mobility to deal with the reduction of distance between productive patches and the long distance
location of raw material sources, which in turn favored exible
operational chains and low technical investment. This study, along
with other recent investigations, stressed that the early Holocene in
the region was a very diverse period with distinct technological and
environmental changes.
On a similar topic, but based on the chemical analysis of organic
remains found on lithic artifacts, N. Mazzia and N. Flegenheimer
identied a variety of resources which were relevant to discuss
paleomobility, paleoenvironmental scenarios and the diversity of
resources consumed in the Pampean region (Argentina) during
the late Pleistocene and the early Holocene. The lithic repertoire
analyzed provides clues on marine resources consumed, suggesting
high mobility that includes different terrestrial and coastal scenarios. Their results stressed the important role of vegetation resources among early hunteregatherers which contrasts with
previous interpretations performed by local archaeologist on high
dependence of animal resources, basically megafauna. This paper,
despite the requirements of good preservation, supports the idea
that the analytic technique of chemical analysis of organic matter
(in this case fatty acids) recovered from lithic artifacts provides
an independent and reliable line of evidence concerning consumed
resources and dietary mobility.
In Patagonia, Argentina, A. Brook, N. Franco, P. Ambrstolo, M.
Mancini, L. Wang, and P Fernandez presented evidence for early occupations in the southern Deseado Massif during the Pleistocene/
Holocene transition. On the basis of sediment and pollen analysis,
the authors suggested that the major human occupations coincided
with wetter conditions. 14C dates conrm that megafauna, more
specically the giant ground sloth, were present after the rst human arrival and became extinct soon afterwards. The lack of evidence of human activities during several early and middle
Holocene periods is explained, at least partially, through the increase of aridity and reduction of water sources during the same intervals inferred from the pollen records. This paper presents new
and interesting evidence on the initial peopling of Patagonia.
Finally, L.A. Borrero presents a very interesting paper on crucial
aspects of the initial peopling process of South America including
the knowledge of the environment by the rst settlers, differential
preservation of archaeological materials and visibility (regional
taphonomy), biogeographic features, and ecological peopling
models, all framed in the theoretical perspective of cultural geography. The author discussed in a detailed form the mechanisms
implied in exploration, colonization, and effective occupation
phases. When this wide range of indicators and perspectives are
applied to the early South American archaeological record, an interesting pattern emerged which indicates that the rst settlers

Guest Editorial / Quaternary International 363 (2015) 1e3

probably were generalist hunteregatherers with high exibility to


exploit different niches, likely following less-resistant routes and
employing a variety of lithic technologies mostly focused on a
wide spectrum of resources, including plants. Finally, Borrero
stressed the importance of evidence emerging from forested landscapes mainly tropical forests from Colombia and Brazil, the key
role of non-utilitarian items and exchange to identify exploration
and/or effective colonization stages, and the importance of empty
lands which reects both low demography during the initial
peopling and a relative knowledge of the local geography, especially during the colonization and effective occupation phases.
The papers included in this volume, covering a wide chronology
and geography, as well as different study materials, approaches and
perspectives, show distinct approximations e from case studies to
broad interpretations e to understand humaneenvironment interactions in America during the Pleistocene/Holocene transition using archaeological indicators along with a variety of evidence.
This reects an emerging pattern regarding the use of multidisciplinary and interdisciplinary approaches to address holistically
the initial peopling of the American continent. This also exemplies
the use of promising evidence and perspectives to enhance our interpretations on the early human history in the Americas, which
the archaeological community will surely nd interesting.
In conclusion, this volume indicates high population dynamism
during the Pleistocene/Holocene transition in America, a crucial
period to understand both the early hunteregatherer dispersal
throughout the continent and their complex interaction with their
surrounding environment. During this time, important adaptations
emerged, including plant cultivation.
We would like to thank the authors for their contributions and
reviewers for their indispensable collaboration. Specially, we

want to thank the Quaternary International editor Norm Catto for


his kind and effective assistance throughout the evaluation and
editing process. M. Delgado and F. Aceituno thanks Carlos Lopez
and Marta Cano for their logistic support to the session above
mentioned and Gustavo Politis for the discussion of several works
here presented. Ted Goebel, Donald Jackson, and Gaspar Morcote
also participated actively in the discussion.
Miguel Delgado*
n Antropologa, Facultad de Ciencias Naturales y
CONICET e Divisio
Museo, Universidad Nacional de La Plata, Paseo del Bosque s/n., 1900
La Plata, Buenos Aires, Argentina
Francisco Javier Aceituno
Grupo Medioambiente y Sociedad, Departamento de Antropologa,
Universidad de Antioquia, Calle 67 No 53-108, AA 1226 Medellin,
Antioquia, Colombia
E-mail address: francisco.aceituno@udea.edu.co.
s Loaiza
Nicola
Grupo Medioambiente y Sociedad, Departamento de Antropologa,
Universidad de Antioquia, Calle 67 No 53-108, AA 1226 Medellin,
Antioquia, Colombia
Temple University, Department of Anthropology, Philadelphia,
PA 19119, USA
E-mail address: nloaiza@temple.edu.
Corresponding author.
E-mail address: medelgado@fcnym.unlp.edu.ar (M. Delgado).
*

Quaternary International xxx (2014) 1e16

Contents lists available at ScienceDirect

Quaternary International
journal homepage: www.elsevier.com/locate/quaint

Paleoindian sites from the Basin of Mexico: Evidence from


stratigraphy, tephrochronology and dating
Silvia Gonzalez a, *, David Huddart a, Isabel Israde -Alcntara b,
Gabriela Domnguez-Vzquez c, James Bischoff d, Nicholas Felstead e
a

School of Natural Sciences and Psychology, Liverpool John Moores University, Byrom Street, Liverpool, Merseyside L3 3AF, UK
Department of Geology and Mineralogy, IIM, Universidad Michoacana de San Nicols de Hidalgo, Morelia, Michoacn, Mexico
Faculty of Biology, Universidad Michoacana de San Nicols de Hidalgo, Morelia, Michoacn, Mexico
d
United States Geological Survey, Menlo Park, CA, USA
e
Department of Geography, Durham University, UK
b
c

a r t i c l e i n f o

a b s t r a c t

Article history:
Available online xxx

We present new data on the stratigraphy, dating and tephrochonology at the most important Paleoindian
sites in the Basin of Mexico. These include: a) Peon Woman III, with the oldest directly radiocarbon
dated human remains (10,755  75 BP); b) Tlapacoya, with two human crania dated to just over 10 ka BP;
c) Tocuila, an important mammoth site with incorporation of fossils and suggested bone tools within the
Upper Toluca Pumice (UTP) lahar (volcanic mudow). The Tocuila site also includes potential evidence
for a layer associated with the Younger Dryas meteorite airburst, with charcoal, iron microspherules,
micro-tektites (melted glass) and volcanic ash, dated to 10,800  50 BP and d) the Santa Isabel Iztapan
mammoths I and II with lithics of Scottsbluff, Lerma and Angostura types and obsidian prismatic blades
but lacking the characteristic uted Clovis type points normally associated with mammoth kills and
butchering and dated after the Pumice with Andesite (PWA) layer between 14,500 BP and 10,900 BP,
before the Younger Dryas interval. These results show that these lithic traditions in Central Mexico are
older than in the Great Plains of USA. Several tephra markers are recognised in the sites that help to
constrain the stratigraphy and dating of the archaeological sequences. However tephra reworking in
marginal lake sites is present and has been carefully considered, especially for the PWA tephra.
2014 Elsevier Ltd and INQUA. All rights reserved.

Keywords:
Late Pleistocene
Mexico
Tephra
Dating
Mammoths
Paleoindians

1. Introduction
The main objectives of this paper are to present and discuss the
complexities of the sedimentology, tephrostratigraphy and dating
related to the most important Paleoindian sites in the Basin of
Mexico: Peon Woman III, Tlapacoya, Tocuila Mammoths and Santa
Isabel Iztapan I and II Mammoths. The overall aim is to investigate
the inuence of volcanic processes and a potential meteorite
airburst event on the development of human and animal populations during the Late Pleistocene/Early Holocene.

* Corresponding author.
E-mail addresses: S.Gonzalez@ljmu.ac.uk (S. Gonzalez), D.Huddart@ljmu.ac.uk
(D. Huddart), isaisrade@gmail.com (I. Israde -Alcntara), gdoguez@yahoo.com.mx
(G. Domnguez-Vzquez), jbischoff@usgs.gov (J. Bischoff), nicholas.felstead@
durham.ac.uk (N. Felstead).

The sites are located at around 2240 m a.s.l. in the at-oored,


Basin of Mexico, which has been a closed hydrographic system
for the past 700 ka. Magnetic polarity stratigraphy does not give
tight constraints but suggests a Brunhes age (Urrutia-Fucugauchi
and Martin del Pozzo, 1993). The basin contained a large lake
approximately 1000 km2 in area, extending within a series of subbasins (Fig. 1), prior to articial recent drainage. The basin limits to
the east are the volcanoes of the northern Sierra Nevada (Tlloc and
Telapon). Pyroclastic ows, ashes, block-and-ash ows, lahars and
uvial sediments make up an extensive volcaniclastic piedmont
linking these volcanoes to the lake basin. Other potential sources of
volcaniclastic sediment are from the monogenetic volcanoes in the
basin and the strato-volcano Nevado de Toluca from which a Plinian
eruption produced the extensive Upper Toluca Pumice (UTP) and
the strato-volcano Popocatpetl, the source of the Pumice with
Andesite (PWA), south-west and south-east respectively (Fig. 1).
The basin climate today is subtropical with summer rainfall predominant (500 and 1000 mm), but around the basin the mountains

http://dx.doi.org/10.1016/j.quaint.2014.03.015
1040-6182/ 2014 Elsevier Ltd and INQUA. All rights reserved.

Please cite this article in press as: Gonzalez, S., et al., Paleoindian sites from the Basin of Mexico: Evidence from stratigraphy, tephrochronology
and dating, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.015

S. Gonzalez et al. / Quaternary International xxx (2014) 1e16

Fig. 1. Basin of Mexico: Location of studied Paleoindian sites and main tephra markers: GBA, Great basaltic-andesitic ash; PWA, Pumice with Andesite and UTP, Upper Toluca
Pumice. The black arrows indicate the dispersion axes for the tephras and the volcanoes that produced them; dashed black arrow inferred dispersion axe for GBA tephra. The sites
are: 1) Peon Hill, 2) Tlapacoya Hill, 3) Metro Man, 4) Chimalhuacan Man, 5) Tocuila Mammoths, 6) Santa Isabel Iztapan Mammoth II.

provide a wide range of temperature, rainfall and vegetation environments. Climatic change in the last 50 ka has driven large
uctuations in vegetation and lacustrine environments in the basin
(e.g. Lamb et al., 2009).

Peon II: Found in June 1957 in the Colonia Peon de los Baos,
comprising a mineralised human skull embedded in travertine,
(Fig. 2a) (Romano, 1964, 1970). Radiocarbon dating of the skull by
the authors was unsuccessful due to lack of collagen.

2. Description and interpretation of studied Paleoindian sites


2.1. Peon Woman III Paleoindian skeleton
The volcano Peon de los Baos was situated within the
former Lake Texcoco and today is located at the northern side of the
International Airport in Mexico City (Fig. 1). Hot spring activity
during the Late Pleistocene to Early Holocene produced layered
travertine, occurring as a conspicuous platform around the volcano. This hill was an important Preceramic locality in the basin,
with at least 4 human skeletons reported to have been found in the
area, 3 embedded in the travertine deposits and one (Peon
Woman III) below the travertine deposits. These skeletons are now
part of the Preceramic human collection at the National Museum of
Anthropology. All these skeletons were found by chance and there
is no information available as to how they were incorporated into
the sediments. They include:
Peon I: Found in 1884 comprising mineralized and fragmented
human ribs and a femur in travertine (Barcena and del Castillo,
1887; Newberry, 1887).

Fig. 2. Peon Hill humans: a) Peon II Skull found in travertine. B) Peon Woman III
skull, oldest directly dated human from the Basin of Mexico.

Please cite this article in press as: Gonzalez, S., et al., Paleoindian sites from the Basin of Mexico: Evidence from stratigraphy, tephrochronology
and dating, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.015

S. Gonzalez et al. / Quaternary International xxx (2014) 1e16

Peon Woman III (Catalogue number: 6-07-1959/DAF INAH):


This is a very important skeleton found in 1959 in Colonia Peon de
los Baos. It is a semi-complete, well preserved, adult female human skeleton (Fig. 2b). The postcranial skeleton was fragmentary
(missing parts included both femurs and tibias and the right perone). The mandible is well preserved showing almost all the teeth
except the left lateral incisor and all the upper and lower teeth
show strong wear. The height was estimated at 1.51 m. Both cranial
and postcranial evidence indicates a female with an age at death
about 25 years. The skeleton was found below a 2 m travertine
sequence in sediments with volcanic tuff characteristics (Aveleyra
de Anda, 1964). The skeleton age was estimated to be Late Pleistocene because it was found below the UTP ash (Mooser and
Gonzlez-Rul, 1961). The deposit yielded no pottery, but threads
of natural bres were attached to the pelvis (still preserved today),
and a polished bird bone and a root fragment with sediments were
collected at the time of the discovery. The skull had very primitive
characteristics (Romano, 1974). Later Gonzalez et al. (2003)
directly dated a sample from the left humerus using AMS C14 to
10,755  75 yr BP, (OxA-10112), conrming a Late Pleistocene age.
(All the dates in this paper are uncalibrated). Detailed descriptions
of the physical anthropology are included in Jimnez-Lpez et al.
(2006).
Peon IV: Fragments of travertine mixed with human bone
including a human skull, found in 1962, with no archaeological
control.

2.1.1. Stratigraphy of Peon Woman III


The stratigraphy for Peon III is re-drawn here from Mooser and
Gonzlez-Rul (1961), including our own eld observations at the
site (Fig. 3). The stratigraphic position of the skeleton is below the
UTP ash layer, derived from a Plinian eruption of Nevado de Toluca
volcano, radiocarbon dated to 10,445  95 yr BP (Arce et al. 2003).
We have also studied sediment samples found attached to the
foot of the skeleton, taken at the time of the discovery, which
consist of a mixture of organic material, molluscs, ostracods, diatoms, phytoliths, quartz crystals and volcanic ash grains, including
small pumice clasts. They represent a saline and alkaline shallow
lake/marsh environment surrounding the island of El Peon at the
time.

2.1.2. Importance
Peon Woman III is the oldest directly dated skeleton from
Central Mexico. The potential association of its pelvis with natural
bres is intriguing and unique to Paleoindian populations of this
age and requires urgent investigation. This Paleoindian woman was
living during the Younger Dryas cold chronozone.
The skull is dolicocephalic, with a cranial index of 70.05, which
reinforces the view that older Paleoindian populations in the
Americas were different from modern Amerindian populations
which are generally brachycephalic, with cranial indices of 80.
Attempts have been made to extract ancient DNA from this skeleton, but despite good bone preservation, with high collagen content (50.3 mg/g), no results have been replicated.
2.2. Tlapacoya
Here we re-assess evidence for early human occupation based
on previous excavations at Tlapacoya Hill (Mirambell, 1967, 1978;
Lorenzo and Mirambell, 1986a) in the SE of the basin (Fig. 1) and
present new stratigraphy, tephrochronology and AMS C14 dates for
the sediment sequence.
2.2.1. Previous work at Tlapacoya
The Tlapacoya site was discovered during motorway construction in the 1960s. Tlapacoya is an andesitic, Miocene volcano which
was at times a peninsula or an island within Lake Chalco depending
on lake levels. Construction work exposed sediment layers that
contained fragmentary animal bones and reddish areas thought to
be associated with burning. This led to an excavation programme at
eighteen localities between 1965 and 1973, concentrated mainly
around the SE hill base, Fig. 4 (Lorenzo and Mirambell, 1986a). The
most signicant evidence for establishing the antiquity of early
human occupation was found in trenches Tlapacoya I, Alpha, Beta
and Tlapacoya XVIII. Excavations in Tlapacoya I produced Late
Pleistocene animal bones (black bear and cervids) in association
with pebbles and what were interpreted to be hearths and artefacts
(Lorenzo and Mirambell, 1986a). Radiocarbon dates of
24,000  4000 BP (A 794b) and 21,700  500 BP (I 4449) were
obtained on humic extracts from charcoal in trench Alpha and one
of 22,000  2600 BP (A 790 A) for a layer containing a supposed
quartz scraper in trench Beta. Tlapacoya II, on the NW hillside,

Fig. 3. Stratigraphic sequence at Peon Woman III (modied from Mooser and Gonzalez Rul, 1961); S Position of the human skeleton.

Please cite this article in press as: Gonzalez, S., et al., Paleoindian sites from the Basin of Mexico: Evidence from stratigraphy, tephrochronology
and dating, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.015

S. Gonzalez et al. / Quaternary International xxx (2014) 1e16

Fig. 4. Tlapacoya Hill: location of previous archaeological trenches (after Lorenzo and Mirambell, 1986a). Also marked with a star is the location of the three new stratigraphic
trenches (A B C) reported here (See Fig. 6a).

produced similar stratigraphy and a swamp cypress trunk (Taxodium mucronatum), with a C14 date of 23,150  950 BP (GX 0959)
from beneath which was recovered an obsidian blade.
A fragmentary human cranium was found in a horizon dated to
9920  250 BP (I 6897) in trench Tlapacoya XVIII (Mirambell,
1986a). No details are given of the precise stratum in which the
cranium was found and no stratigraphic section is reported for this
trench. However, Garca-Brcena (1986) lists the date as obtained
on unspecied material from the base of the UTP. The absence of
stratigraphy for Tlapacoya XVIII and the lack of details related to the
craniums recovery make this nd frustrating for later researchers.
In 1968, archaeologists were informed of an earlier discovery of
another human cranium that had been re-buried. The original
location of the cranium lay 50 m to the north of trench Beta
(Mirambell, 1986a). Mirambell points out that the specimen shows
features unlike those generally encountered in Mexican crania, and
Romano (1974) listed the cranium as dolicocephalic (cranial index
of 67.17). This specimen has now been directly AMS radiocarbon
dated to 10,200  65 BP (OxA-10225) and is known as Tlapacoya I
skull, Fig. 5 (Gonzalez et al. 2003).
Lithic materials found in trenches Alpha, Beta and Tlapacoya 1
were reported as evidence of very early human activity on a lake
beach by Lorenzo and Mirambell (1986a). Other interpretations are
more plausible. The 14C dates for these levels suggest human activity considerably older than the age of the cranium from trench
XVIII. The suggested artefacts provide the cultural evidence based
on a lithic assemblage composed of 2500 andesite akes. The akes
are of the same local bedrock lithology that fractures naturally into
very sharp akes. They occur abundantly on todays surface. There
were three obsidian akes and two small bone fragments, claimed
to be worked. From a poor photo (Mirambell, 1986b), the bone
fragments appear to be a distal second phalanx from an ungulate
and a small ake of longbone. The phalanx was suggested by Mirambell to be a whistle and the ake as a point but both are probably
simply broken bones, selected from the total sample of over 100
bones in the vicinity of the hearths in Tlapacoya I Alpha. Most of
these bones appear to be cervids and a single black bear (Ursus
americanus), along with a small number of coyote (Canis latrans)

and racoon (Procyon lotor). Alvarez (1986) emphasises the complete


absence of any traces of cut marks, or deliberate breakage of the
bones.
The interpretation of Tlapacoya Alpha sediments as a lake beach
depends on assumptions about former lake levels and on the
deposition of what were described as smooth stones, pebbles and
gravels representing a beach (Mirambell, 1978). Many of the deposits in all of the Tlapacoya trenches were recognised as volcanic,

Fig. 5. Tlapacoya I (10-1961-62-DAF/INAH),


10,200  65 yrs BP, scale in cm.

radiocarbon

dated

crania,

Age:

Please cite this article in press as: Gonzalez, S., et al., Paleoindian sites from the Basin of Mexico: Evidence from stratigraphy, tephrochronology
and dating, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.015

S. Gonzalez et al. / Quaternary International xxx (2014) 1e16

Fig. 6. a Tlapacoya, location of studied stratigraphic trenches (A B C) in a prole from the hill side to the lake. b Tlapacoya new trenches stratigraphy, showing the position of
samples taken and with black stars the position of the new AMS radiocarbon dates obtained in organic materials (See Table 1).

mainly tephra falls into the lake, interbedded with diatomites.


Diatom recovery and interpretation was less than optimum
(Bradbury, 1986, 1989) and species from deposits associated with,
or close to, the supposed beach hearths were those of brackish

conditions (Bradbury, 1971). There were occasional lake high


stands, but much of the diatom accumulation took place in smaller,
alkaline pools. This impression is derived from paleoenvironmental
interpretations from the former Lake Chalco deposits (Urrutia-

Please cite this article in press as: Gonzalez, S., et al., Paleoindian sites from the Basin of Mexico: Evidence from stratigraphy, tephrochronology
and dating, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.015

S. Gonzalez et al. / Quaternary International xxx (2014) 1e16

Fucugauchi et al., 1995; Caballero Miranda, 1997) that characterise


the Late Pleistocene lake as variable in depth, oscillating between
fresh and saline water. As for the suggested beach deposits, it is
hard to see how even full lake levels of around 20 m would produce
wave action to form beach pebbles and these pebbles would be
rounded. It is more likely that the coarse, angular gravel layers
found in the trenches represent weathered scree carried down the
Tlapacoya Hill slopes to accumulate at the slope-break below the
steep face. The suggestion of rounded edges on the andesite
akes is unsubstantiated by shape analyses in Mirambell (1978),
and doubted by Gonzalez and Huddart (2008).
The Tlapacoya environment during the formation of these deposits was not the pleasant lakeside locality as represented by
Lorenzo and Mirambell (1986a). Much of the Late Pleistocene in the
area was interrupted by frequent and often heavy ash falls, as major
volcanic eruptions took place. Despite the obvious impact of such
major volcanic eruptions on the environment and human and animal populations, this caused little comment from the archaeologists at the time of the excavations. Hence there is uncertainty from
the previous work as to the stratigraphy, its environmental interpretation, and the true age of any human presence. We have
therefore taken the opportunity to re-investigate the Tlapacoya
stratigraphy, rene its dating by obtaining new AMS radiocarbon
dates, and analyse the volcanic ashes. Three new trenches were
excavated to re-study the Tlapacoya sequence in detail, and are
reported here.
2.2.2. New excavated trenches at Tlapacoya
Three new stratigraphic trenches were excavated (A, B, C) by the
authors (Fig. 6a and b). The aim was to locate them as close to those
excavated in the SE part of the hill (Trenches Alpha and Beta) reported by Lorenzo and Mirambell (1986a). This proved impossible
because of the large population and settlement growth in the area
that has completely altered the appearance from that shown in the
photos in Lorenzo and Mirambell (1986a). Ironically, whilst settlement growth has reduced the choice of possible excavation areas
it did provide a ready-made trench in undisturbed deposits from a
proposed septic tank, in approximately the same position as Tlapacoya I. This is designated as trench Lake Margin B in Fig. 6b. A
roadway cut into the hillslope exposed trench Hillside A and the
third trench, Lake C was excavated in open ground midway between the trenches Tlapacoya XVIII and Tlapacoya I of the original
excavations.
2.2.3. Tlapacoya new stratigraphy and tephrochronology studies
The motorway construction removed the top Holocene
sequence around the SE of Tlapacoya Hill. The new excavated
stratigraphy (Fig. 6b) is complex and consists of diatomites, lake
sediments, ash falls, reworked tephra, clays, organic-rich sediment,
and slope deposits. No bones, hearths, or artefacts were recovered.
In the cut in Hillside A, the sequence is dominated by volcanic
ashes, some in situ (e.g. samples A5 and A2), whilst others were
reworked downslope. The coarse gravel unit towards the base is
composed of sub-angular/angular rock fragments, up to 20 cm b

axis, embedded in grey/beige, sandy ash. Ashes A2, A3, A5 contain


white pumice up to 5 cm b axis. The unit A6 consists of topsoil
and angular gravel up to 10 cm b axis, with pumice fragments in a
sandy matrix.
The basal organic deposits in trench Lake margin B have been
radiocarbon dated from 22,610  100 BP to 20,960  130 BP (See
Table 1) and consist mainly of peat, alternating with diatomites and
silts/clays, all associated with a shallow, freshwater lake. The
middle of the section is dominated by diatomites, but B11 is a greyblack ash, with thin layers of brown silt and diatomite, whilst B15/
16 are grey/black ashes with some diatoms. Layers B18/19 are
composed of a rounded to sub-angular gravel, up to 30 cm b axis,
embedded in a dark grey, sandy ash matrix, with few diatoms and
pumice fragments. Above is a grey, sandy-silty ash (B20), with
isolated lenses of black ash (B21). The sequence is capped by grey,
sandy silts, with pumice lenses up to 10 cm across and gravel
fragments, with some animal burrows.
The basal unit in Trench C is a black peat, above 2 cm silty, grey
ash. The peat was radiocarbon dated to 13,030  70 BP (See
Table 1). Above are thin ashes, some with pumice, and all showing
diatoms and one white, silty diatomite with charcoal fragments.
The thick unit above (C6, C7 and C8) is composed of gravels and
sands, matrix-supported, occasionally organic-rich, with some
caliche lenses, sub-rounded and rounded pumice, up to 3.5 cm b
axis and angular rock fragments, up to 8 cm b axis. In C6 there are
diatoms and wood fragments. Units C9, C10 and C11 are units of
white/grey, silty-sandy ash, with few diatoms. The uppermost layer
(C12) is a sandy-silty, brown soil, with many roots and pumice
fragments, up to 7 mm b axis.
Many of the volcanic ashes found were mixed and reworked
with lake sediments. The only way to study their true chemical
composition was by obtaining their major oxide geochemistry in
single grain microscopic tephra, using a Cameca SX100 Electron
Probe Microanalyser. The tephra major element geochemistry obtained is given in the Supplementary Data and summarised in
Table 2.
2.2.4. Interpretation of the new Tlapacoya stratigraphy
2.2.4.1. Trench hillside A. The sequence is dominated by ashes that
have been reworked downslope, although A5 and A2 are in situ, and
depending on the supply source there is usually a mixture of PWA
(59 to 63% SiO2) and UTP tephra (w69.9% Si02). In the lowest
sample A0 there are again two tephra populations, one with a mean
of 61% Si02 (PWA) and the other with a rhyolitic composition (mean
of 73% Si02) which is considered to be reworked from a tephra from
Tlaloc volcano. There is similar reworking of tephra populations
from Lake Texcoco marginal areas in the Tepexpan Paleoindian and
Tocuila Mammoth sites (Gonzalez and Huddart, 2007; Lamb et al.,
2009; Gonzalez et al., 2014). The 10 cm unit of sub-angular
andesite gravels in an ash matrix is interpreted as reworked hillside scree, incorporated with ash by hillslope processes, such as
insolation weathering and then slopewash, or soliuction and is not
a beach as interpreted by Lorenzo and Mirambell (1986a).

Table 1
Tlapacoya Uncalibrated AMS C14 results in organic materials. (See Fig. 6b).
Sample

Laboratory
number

Description

14

Tlapacoya I
Trench C, Layer C0
Trench B: Layer B8
Trench B: Layer B6
Trench B: Layer B3

OxA-10225
Beta-131855
Beta-131856
Beta-131857
Beta-131854

Human cranium
Burned vegetation
Peat
Peat
Wood fragment

10,200
13,030
20,960
21,800
22,610

C BP

Type of
analysis






65
70
130
190
100

AMS
Standard
Standard
Standard
AMS

d13C
16.1
25.0
25.0
25.0
27.0

Please cite this article in press as: Gonzalez, S., et al., Paleoindian sites from the Basin of Mexico: Evidence from stratigraphy, tephrochronology
and dating, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.015

S. Gonzalez et al. / Quaternary International xxx (2014) 1e16


Table 2
Summary of Geochemistry of Tlapacoya Tephra samples (see Fig. 6b).
Tlapacoya
Trench A
A0
A0
A1
A1
A2
A3
A3
A4
A5
A5
Trench B
B1
B1
B3
B3
B11
B16
B16
B16
B21
B21
Trench C
C2
C2
C5

Mean SiO2 %

S.D.

Variance

Range

5
3
6
8
11
9
3
12
14
5

61.0
73.41
62.25
69.36
59.07
59.40
69.39
63.09
61.46
69.99

3.90
0.33
2.94
3.37
2.17
2.76
3.38
1.70
2.17
1.96

15.2
0.11
8.63
11.37
4.71
7.64
11.44
2.9
4.71
3.86

56.75e65.44
73.13e73.78
58.69e64.87
66.35e75.28
56.73e60.89
56.12e63.91
66.70e73.19
60.05e65.53
58.60e65.73
68.67e73.24

7
5
4
6
11
4
1
1
10
1

60.14
69.83
70.78
61.16
59.88
71.50
54.52
64.3
57.39
66.15

3.27
2.36
1.47
3.95
2.79
0.73

10.67
5.58
2.16
15.6
7.78
0.54

54.16e63.75
67.50e73.66
69.21e72.75
56.90e65.16
53.62e63.17
70.69e72.15

0.53

0.29

56.58e58.25

6
4
15

60.95
66.93
65.08

3.32
0.57
2.38

11.03
0.33
5.67

56.43e65.49
66.33e67.66
58.43e68.78

2.2.4.2. Trench Lake side B. The lowest 2.50 m of this section is


dominated by marginal lake sediments (diatomites, silts and silty
sands) and thin organic layers, the lowest 85 cm dated by radiocarbon dating between 22,610  100 and 20,960  130 BP. There are
thin volcanic ashes interlayered throughout this sequence. Samples
B1 and B3, associated with the oldest radiocarbon date obtained
from wood, have two tephra populations with mean values between 60 and 61% Si02 (andesitic) and 69 to 71% SiO2 (rhyolitic).
Probably, the andesitic composition is derived from an eruption of
Popocatpetl Volcano that is known at this period. Tephra TB11 is
14 cm thick and has only one tephra population, with a mean of 59%
Si02 and is interpreted here as the PWA deposited as airfall into the
lake. In the basin it has a thickness range from 5 to 30 cm, and can
be traced to Popocatpetl, with an age of around 14,500 BP (Mooser,
1967; Garca-Brcena, 1986; Ortega-Guerrero and Newton, 1998).
This is followed by lake sediments (diatomites and silts) below the
22 cm thick gravel unit in a sandy ash matrix, with pumices and
diatoms (samples B18 and B19). As in trench Hillside A, this unit is
interpreted as a slopewash deposit into the marginal lake, or close
to the shoreline from Tlapacoya Hill. It cannot be a lake beach
gravel, as the gravels are too angular and there is inadequate fetch
to produce a beach gravel. Directly above this unit is a 2e3 cm
irregular, organic horizon, interpreted as a subaerial palaeosol,
capped by a cream, massive, silty-sandy ash, with pumice and
gravel clasts up to 5 cm b axis. Within this unit are lenses of black
ash up to 2 cm thick. Sample B21 from this latter unit is dominated
by a basaltic-andesitic geochemistry (Si02 of 57%). This tephra is
likely to have been derived from a local monogenetic cinder cone
from the Santa Catarina Range.
2.2.4.3. Trench Lake C. This sequence is constrained by a radiocarbon date of 13,030  70 BP from the base of the trench, unit C0,
which is made of charred vegetation. Above there is a thin sequence
of coarse pumice C1, with a bulk chemical composition of 59% Si02,
C2 is a silty-sandy black ash, with many diatoms and thin diatomites. C2 is interpreted as marginal lake reworking of PWA,
along with an unknown tephra. For example C2 has two mean
tephra populations of 60.9% Si02, interpreted as PWA and 66% SiO2.

C5 has only one population, with mean SiO2 of 65%. The thickest
unit in this trench is the 1.03 m, hillslope deposit dominated by
reworked PWA pumice and ash. C7 has a tephra population with a
mean Si02 of 63%, is a gravelly matrix-supported sand, with rock
fragments between 5 and 8 cm baxis, some caliche lenses, diatoms, sh scales, and occasional wood fragments. It is capped by a
13 cm palaeosol. The top of the sequence is dominated by 30 cm
white to grey, tripartite, silty-sandy ash, (mean Si02 of 69.92%),
interpreted as the UTP in situ. This unit is important because it is in
this layer that the stratied human cranium was excavated from
Trench XV111 (Lorenzo and Mirambell, 1986a). This tephra at Tlapacoya was C14 dated to 9920  250 BP (I-6897), for the base of the
ash sequence, correlating more or less with the AMS date obtained
directly for the unstratied human cranium.
2.2.5. Tlapacoya radiocarbon dates
Samples for C14 dating were taken from animal bones obtained
from the original excavations described in Lorenzo and Mirambell,
(1986a). Two phalanges of Ursus americanus (DP-958 and DP-957)
were selected because of their association with the hearth in
Trench Alpha. As we were unable to locate the fragmentary in situ
human cranium from trench XVIII, the third sample was taken from
the second, unstratied human cranium, number 10-1961-DAF/
INAH. Unfortunately neither bear phalange gave a date, because of
poor collagen preservation, but the cranium was dated to
10,200  65 BP (OxA-10225). This is one of the oldest directly dated
humans currently known from Central Mexico (Gonzalez et al.
2003). Organic materials were selected from the new trenches
and AMS radiocarbon dated (Table 1), providing reliable dates (with
low error values) and those from trench Lake margin B are in
stratigraphic order.
2.2.6. Tlapacoya I Human Cranium, 10-1961-62-DAF/INAH
The dated unstratied cranium (Fig. 5) is well preserved, with
collagen content of 10.9 mg/g, but has lost all of its facial and palatal
bones, together with the basilar and condylar parts of the occipital.
A portion of the left squamous part of the occipital was removed for
the radiocarbon AMS determination. The cranium is dolicocephalic
(Gonzalez et al. 2003) and the parietal eminences are welldeveloped. On the basis of size, robusticity, and sutural fusion, it
appears to be a mature, adult male, with a marked glabella and well
developed supraorbital tori, external occipital region and mastoid
process. The occipital region is prominent and has several wormian
bones, while the frontal bone is centrally ridged and bears a small
depression near bregma that may have been produced by a trauma
earlier in life. Further possible evidence of early trauma is seen in
slight depressions on either side of the sagital suture close to
lamda, and only the right parietal foramen is present in this area.
The right orbit has a supraorbital notch whilst the left has two
foramena, the more medial one larger than the lateral. Slight pitting
(criba orbitalia) may be seen in the frontal bone surface within the
orbits, a condition indicative of anaemia (Roberts and Manchester,
1997), but there is no evidence of pathology of the tempromandibular articulation in either the mandibular fosase, or the
articulareminences. There is no obvious evidence of deliberate
perimortem, or immediately postmortem activity, such as cutting.
The endocrania is covered by a sediment layer, and it has not been
possible to describe its morphology.
2.2.7. Importance of Tlapacoya
The claims for early human occupation at this site from supposed hearths associated with Pleistocene vertebrate bones and
obsidian blades on beach gravels remain controversial, and we did
not nd any evidence to support the claims. However the two
fragmentary human skulls, found during the original excavations,

Please cite this article in press as: Gonzalez, S., et al., Paleoindian sites from the Basin of Mexico: Evidence from stratigraphy, tephrochronology
and dating, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.015

S. Gonzalez et al. / Quaternary International xxx (2014) 1e16

have now been dated, one directly by radiocarbon AMS and the
other indirectly by stratigraphy. It is unfortunate that they lack
associated artefacts. The fact that one of the human skulls was
found associated with the UTP tephra is important. We have found
other Paleoindian human remains associated with this volcanic
layer (Metro Man and Chimalhuacan Man), that conrms that there
was a Late Pleistocene human population living on the shores of
Lake Chalco at the time of this major volcanic eruption during the
Younger Dryas period, with different cranial morphology to Modern Amerindians. These human skeletons associated with ash from
the eruption, were buried quickly, which allowed their preservation. We do not know if the eruption killed these humans.
We have found that by studying the detailed geochemistry of
individual volcanic glass in the Tlapacoya sequence and other
Paleoindian sites in the basin, we were able to recognise reworking
and mixing of the ash layers. We found no sedimentological evidence for the presence of coarse gravel, Late Pleistocene beaches
as interpreted by Lorenzo and Mirambell, (1986a) and we propose
that instead they are angular scree deposits derived from the
hillslopes.
3. Metro Man and Chimalhuacan Man Tephra dating
Commonly, human and animal bones of presumed Late Pleistocene age found around the lake sites in the basin and other
Central Mexico lakes are highly mineralised, heavy, and display a
distinctive black coloration. Examples include the Metro Man and
Chimalhuacan Man skulls (Fig. 7). When we attempted direct AMS
radiocarbon dating on them, we found that there was almost no
collagen preserved, making C14 dating impossible. During detailed
examination inside these two human crania, sediment samples
were collected and analysed using electron microprobe studies.
Their major element geochemistry was then compared against
known volcanic ashes that were deposited during the Late
Pleistocene to Early Holocene in the Basin to try to date the
specimens using tephrochronology.
The Metro Man skull was found in 1970 at 3.10 m depth, during
construction of the Metro Balderas station in Mexico City centre,
embedded in the UTP ash (Mooser, 1967). Volcanic ash taken from
the skull gave results of 63 to 71.2% SiO2 showing some mixing
(Table 4), but mainly associated with the UTP tephra.
The Chimalhuacan Man skeleton was found in 1984, in Colonia
Embarcadero, Mexico State (Pompa y Padilla, 1988) but there are no

Fig. 7. a. Metro Man skull showing intense mineralisation. Fig. 7b. Chimalhuacan Man
skull. Both skulls were dated indirectly using tephrochronology methods, see the text.

published records of the stratigraphic context. Sediment from inside the skull was a mixture of lake sediments, diatoms and volcanic ash. The electron microprobe analysis of the ash indicates
SiO2 values between 62 and 77%, indicating a mixture of ashes, with
the latest probably corresponding also to the UTP eruption
(Table 4). For this reason, we interpret that the skull has an indirect
date of around 10,500 BP and not 33,000 BP as suggested previously
using obsidian hydration dating (Pompa y Padilla, 1988).
4. Tocuila Mammoths Site
Tocuila is located in a western Texcoco suburb, in sediments of a
former near-shore, Late Pleistocene, higher level of Lake Texcoco
(Fig. 1). A review of detailed recent stratigraphic work and a reinterpretation of the sedimentology, dating and origin of the
Tocuila Late Pleistocene mammoths site has been reported by
Morett et al. (1998) and more recently by Gonzalez et al. (2014). The
main objectives were to understand the stratigraphic complexities
and tephra sequence at the site and to describe a newly recognised
layer that may represent the suggested Younger Dryas meteorite
impact event (Israde et al. 2012). This layer is between 6 and 10 cm
thick, and consists of a mixture of lake-fall volcanic ash, magnetic
Fe microspherules, micro-tektites (melted glass), charcoal and diatoms. If the interpretation is correct, the occurrence extends the
geographic range of this Younger Dryas event to the Basin of
Mexico. Here, we present a short summary of the results and
conclusions derived from this recent work because of the importance and relevance of the site for the interpretation and understanding of the paleoenvironment in the Basin of Mexico during the
Late Pleistocene to Early Holocene transition.
4.1. Tocuila: previous work
In 1996, the remains of at least seven mammoths (Mammathus
columbi) were excavated, see Fig. 8 (Morett et al., 1998) and evidence for the presence of worked mammoth bones in the bone
assemblage was presented (Arroyo Cabrales et al. 2001). Subsequently, sedimentological and tephrochronological work from the
deposits was reported by Siebe et al. (1999) and Gonzalez and
Huddart (2007).
4.2. New stratigraphic work at Tocuila
The stratigraphic sequence at Tocuila is complicated (Fig. 9), but
it can be explained in terms of a lake marginal sequence that is cut
by a channel inlled with lahar (volcanic mudow) sediments
(Gonzalez et al. 2014). It is in the channel inll where the mammoth
bones were found, embedded in the lahar. The stratigraphy in Fig. 9
shows the channel wall cut in the marginal lake sequence, with the
basal sequence composed of grey clays and a black, indurated,
sandy basaltic andesitic ash.
The channel base is eroded into a 55 cm layer of brown clay,
with root casts. Incorporated into this unit there are lenses
(4 cm  5 cm) of sandy ash (Sample A-5a), which shows two tephra
populations. One has a mean SiO2 of 55% and the other a mean SiO2
of 74%. This is unconformably overlain by a conspicuous unit, 6e
10 cm, of laminated, black, reworked silty ash, waterlain, with large
amounts of charcoal, micro-tektites (siliceous glass spherules), diatoms and magnetic iron spherules (see location of sample Toc-6 in
Fig. 9). This layer shows two populations of tephra shards (means of
55% and 67% SiO2). This is overlain by 45 cm of sandy silt, with root
casts and diatoms. There are also pumice clasts, up to 1.0 cm in
diameter in this silt, which grades transitionally into a further silt
unit, with pumice clasts and root casts. This is overlain by silty clay,
with no root casts in contrast to the units below. It is followed by

Please cite this article in press as: Gonzalez, S., et al., Paleoindian sites from the Basin of Mexico: Evidence from stratigraphy, tephrochronology
and dating, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.015

S. Gonzalez et al. / Quaternary International xxx (2014) 1e16

Fig. 8. Photograph showing the Tocuila mammoths trench in 2013.

Fig. 9. Tocuila mammoths general stratigraphy in the main mammoth trench (Museum), after Gonzalez et al. (2014).

Please cite this article in press as: Gonzalez, S., et al., Paleoindian sites from the Basin of Mexico: Evidence from stratigraphy, tephrochronology
and dating, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.015

10

S. Gonzalez et al. / Quaternary International xxx (2014) 1e16

9 cm of pale-grey, silty ash, which becomes lighter coloured towards the top of the unit. This unit has two shard populations with
mean SiO2 values of 66% and 74%. This grey-white ash is overlain by
8 cm of dark grey, laminated, silty clay with diatoms. The sequence
is topped by 45 cm of sandy clay and grey-brown, silty clay, with
gastropod shells throughout.
The lahar channel sequence is dominated by an ungraded,
poorly sorted unit, 1.75 m thick which has a silty-sand, ash matrix.
It includes rounded silt balls up to 1 cm in diameter, pumice lapilli
clasts, up to 3.5 cm in diameter and andesitic lithic clasts. The clasts
show a macrofabric indicating SE-NW transport. This unit contains
charcoal fragments throughout, which have been dated by AMS
radiocarbon dating. Towards the base of this unit there is a concentration of over 1000 animal bones. Over 90% belong to Mammuthus columbi, including three almost complete skulls, two
incomplete skulls, and four mandibles (Morett et al. 1998). The
skeletons are disarticulated and elongated bones tend to be subhorizontal and aligned in the same direction as the clasts, suggesting ow alignment. Excavation has yielded other animal
species, including fragments of horse, bison, duck, goose, amingo,
rabbit, camel, turtle, and tortoise (Morett et al. 1998).
The erosional channel margin shows small lenses of reworked
PWA ash. The lahar deposits are composed of four units (Fig. 9);
units 1, 2, and 3 have small, basal, erosional channels that are
inlled with convoluted, brown, clayey silt.
4.3. Interpretation of the Tocuila stratigraphy
The overall sequence is interpreted as a transition from deeper
water lake to a more nearshore facies. The lowest black basalticandesitic ash is laminated, has diatoms and it is interbedded with
lake sediments. It is correlated with the Great Basaltic Ash (GBA) or
Tlahuac ash, dated by Mooser (1997) to 28,600  200 BP.
Overlying the GBA are uniform, lacustrine, grey clays, with white
root casts, although in between there is one conspicuous dark grey

tephra unit 6 to 10 cm thick (Sample Toc-6) that has a mixed


andesitic and rhyolitic ash compositions. In this layer are microscopic magnetic iron spherules, tektites (melted glass) and large
amounts of charcoal (Fig. 10), dated by AMS radiocarbon to
10,800  50 BP, contemporaneous with the Younger Dryas
Boundary Layer (YDB). This unit possibly contains material from the
controversial YDB Meteorite Airburst layer, but further detailed
geochemical work is now underway to test this hypothesis. The YD
Meteorite Airbust Layer has been recognised in other important
mammoth/Paleoindian sites in SW America, e.g. Murray Springs,
Blackwater Draw, Topper, Arlington Springs, and it is associated
there with black organic mats, nanodiamonds and elevated iridium
concentrations in bulk sediments (Firestone et al. 2007; Kennett
et al. 2009). It has been suggested that this impact is associated
with the megafuna extinction and the onset of the Younger Dryas
cold interval (Firestone et al. 2007). However the interpretation is
still controversial and there are other possible interpretations for
the presence of large amounts of magnetic Fe microspherules, such
as derived from spores or from volcanic processes. This layer has
been identied in Lake Cuitzeo sediments in Central Mexico (Israde
et al. 2012) with Fe microspheres and nanodiamonds, but Tocuila
could be potentially the rst site in the Basin of Mexico. However
we emphasize that the Tocuila bone assemblage is not the result of
the meteorite airburst event: the mammoth bones were incorporated into the lahar sequence produced after the UTP ash deposition, after w10,500 BP (see discussion below).
On top of the lacustrine clays there is a thicker, coarse ash,
composed of subrounded pumice and lithics with a mixed andesitic
and rhyolitic tephra composition and it is interpreted as reworked
PWA tephra. Above this reworked PWA tephra there is another
lacustrine sequence of clays or silts, with root casts.
Outside the main mammoth trench above this lake sequence
was found a series of three, in situ, rhyolitic ashes, correlated with
the UTP tephra (Fig. 1), dated by Bloomeld and Valastro (1977) to
about 11,600 BP. However, Arce et al. (2003) dated this tephra to

Fig. 10. Location of possible Meteorite airburst layer in the channel wall at Tocuila. a) Example of typical microscopic Fe magnetic spherule, scale 30 microns (Sample Toc- 6); b)
Chemical analysis from the Fe micro-spherule using the scanning electron microscope.

Please cite this article in press as: Gonzalez, S., et al., Paleoindian sites from the Basin of Mexico: Evidence from stratigraphy, tephrochronology
and dating, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.015

S. Gonzalez et al. / Quaternary International xxx (2014) 1e16

11

Fig. 11. Lithics from the Santa Isabel Iztapan Mammoths. 11a) Lithics from Santa Isabel Iztapan Mammoth I, after Wormington, 1957: 1. Dark grey, int projectile point with a white
patina. 60 mm long plus 1e15 mm because of the break, 27 mm wide; thin and delicate due to ne aking, no medial ridge. Extremely ne pressure retouching along the edges.
Conforms to general features of Scottsbluff type in shape, proportions and aking technique, but much thinner. In Great Plains always less than 8000 BP and associated with fossil
bison. 2. Two edged scraper of black obsidian, with chipped parallel edges worked by ne, although irregular, pressure retouching. 36 mm long and maximum width 27 mm 3. Flake
obsidian knife, 50 mm long and 24 mm wide, one edge crudely pressure retouched to form a scraping edge in a wide arc whilst the other bears three concentric arcs, the central one
being the deepest serving as a spokeshave. 4. Fine-grained grey int, roughly triangular in outline. One face pressure aked into a steep bevel and classied as an end scraper,
43 mm long and 35 mm wide. 5. Prismatic obsidian ake knife, 59 mm long and 17 mm maximum width. Both edges pressure-retouched. 6. Clear, grey int blade, 54 mm long and
19 mm maximum width, extremely ne marginal retouching. 11b) Lithics from Santa Isabel Iztapan Mammoth II, after Wormington, 1957: 1. A red dacitic/andesitic, lanceolate point
without shoulders, symmetrical shape, the base slightly concave; length 80.2 mm, width 27.4 mm and thickness 8.5 mm. Fine pressure aking over all the basal edges. Originally
dened as an Angostura point by Aveleyra de Anda (1955), although Wormington (1957) disagrees with this designation. 2. Brown int, leaf-shaped projectile point; lacks the distal
extremity. The chipping is bifacial, with scars of irregular akes over both sides and ne pressure aking along the edges, maximum length 61.3 mm, maximum width 24.4 mm,
maximum thickness 8.1 mm 3. Light coloured, chert, biface knife; percussion completely over both faces and secondary retouching on some small sectors of the edges. Probably
originally had a laurel-leaf form; length 67.2 mm, width 34.9 mm and 9.3 mm thick.

around 10,500 BP. The thickness in Tocuila of 44 cm is slightly


thicker than the reported UTP deposits at Tlapacoya between 30
and 37 cm (Gonzalez et al. 2001; Gonzalez and Huddart, 2008).
4.4. Age of the Tocuila deposits: radiocarbon dating
The lahar deposits in the Mammoth Trench have been previously radiocarbon dated by Siebe et al. (1999); Morett et al. (1998b)
and Arroyo Cabrales et al. (2003). The value of these dates has been
discussed by Gonzalez and Huddart (2007) as they are not in
stratigraphic sequence. It is not surprising to nd such variability in
a lahar deposit that is by its nature a reworked deposit. An average
date of about 11,188 BP has nevertheless been given for this deposit
by Morett et al. (1998) and Arroyo Cabrales et al. (2003), but we do
not consider this procedure appropriate because the deposit is
mixed. From our stratigraphic observations and C14 results we
interpret the lahar sequence at Tocuila with the mammoth remains
to be associated with the deposition and rapid reworking of the UTP
tephra shortly after 10,500 BP.
4.5. The Tocuila bone assemblage interpretation
The excavated area in the lahar has produced approximately one
thousand bones, mostly of the Columbian mammoth (Mammuthus
columbi). These represent at least seven individuals ranging from
young to adults. The skeletons were not complete, but some show
articulation indicating that they have not been transported far. It
has been argued by Arroyo-Cabrales et al. (2001) and Johnson et al.
(2001) that the presence of humans is indicated by dynamic impact
fracturing on mammoth long bone segments and fracture debris.
The small assemblage includes a bone core with a prepared platform and scars from the removal of a number of large cortical akes
and a cortical bone ake with remnant platform preparation. The
cortical ake conjoins with the central ake scar on the bone core.

The assemblage is interpreted as mammoth bone quarrying to


produce cores for transport elsewhere. This is not a subsistence
activity but a technological one aimed at securing raw material
shaped into a transportable, useable form. It could be an activity
occurring together with butchering of a mammoth, or an independent activity but is a specialised activity requiring fresh
mammoth bone.
How the mammoth bones became incorporated into the lahar
sediments has been debated: A) were they previously deposited in
the channel either as the result of attritional accumulation or a
catastrophic event (meteorite burst?) and covered subsequently by
the lahar? or B) were the bones transported into the channel with
the lahar? It seems unlikely that the UTP lahars killed the mammoths, because two direct AMS dates on their bones gave dates of
11,100  80 BP and 11,255  75 BP, indicating that they were
already dead when incorporated into the UTP distal lahars and
concentrated in the lahar channels in the lake nearshore (Morett
et al. 1998; Siebe et al. 1999). The time gap between the mammoths death and their incorporation into the lahar means that the
skeletons were lying around the lake shore for several hundred
years. The fact that bone quarrying requires fresh bone means that
humans killed, or scavenged the bones for tool production well
before the skeletons and bone tools were incorporated into the UTP
lahars.
5. Santa Isabel Iztapan Mammoths
5.1. Site description
Santa Isabel Iztapan (Figs. 1 and 11) has two of the most
important mammoth sites because of the indisputable association
with lithic points (Aveleyra Arroyo de Anda and MaldonadoKoerdell, 1952, 1953; Martinez del Rio, 1952; Aveleyra Arroyo de
Anda, 1955, 1956; Wormington, 1957; Gonzalez et al. 2006). The

Please cite this article in press as: Gonzalez, S., et al., Paleoindian sites from the Basin of Mexico: Evidence from stratigraphy, tephrochronology
and dating, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.015

12

S. Gonzalez et al. / Quaternary International xxx (2014) 1e16

mammoths in the Basin of Mexico have been described by Reyes


(1923), Lorenzo and Mirambell (1986b), Carballal-Staedtler (1997)
and Gonzalez et al. (2006) and most are found around the marginal
fringes of the former Lake Texcoco. Into this former lake a series of
currently ephemeral steams fed as part of the low-angled delta of
the Rio San Juan in the Santa Isabel and Tepexpan area, but the lake
at the time of the deposition of the mammoth bones must have
been no more than 70 cm in depth.
The lithics found associated with the two mammoths are
interesting because they are not uted, or Clovis type (Figs. 11a and
b). Haynes (1969) discussed three important mammoth sites in
North America with associated artefacts that are typologically not
Clovis, that included the Santa Isabal Iztapan mammoths lithics,
indicating that there was no date for the nds. The uncertainty
regarding the typology and dating of lithics at Santa Isabel Iztapan
was also mentioned by Tankersley (2002). This is not exactly true
because Mooser and Gonzlez-Rul (1961) interpreted the mammoths stratigraphic position as below the PWA ash at both Santa
Isabel Iztapan I and II (Fig. 12).
As Mooser (1967) dated peat under the PWA with an age of
14,700  280 BP and Lozano-Garca et al. (1993) also published a
second date of 14,450  100 BP it was possible that the mammoths
were killed around these Pre-Clovis dates. Pichardo (2000)
concluded that it was imperative that a detailed re-examination
of the stratigraphy of the Tepexpan, Iztapan and Peon sites be
made to conrm the tephra sequence proposed by Mooser and
Gonzalez-Rul in 1961. In a new trench excavated as close as
possible to the same location of the Santa Isabel Iztapan II site,
Gonzalez et al. (2006) reported that even now the real age of the
mammoth is not very clear, but further tephrochronology studies
are underway to clarify the chemical composition of the volcanic
ashes found in the sequence. This new work is reported here to
clarify further the age and signicance of this site and its associated
mammoth remains and lithics. The lithic assemblage lacks uting
(Figs. 11a and b) and includes Scottsbluff, Lerma (laurel-leaf), and
Angostura type points that have been reported in the USA associated with bison bones. For this reason they are thought to be
younger than Clovis points (Wormington, 1957). The Mexican
assemblage has been given dates from older than the PWA
(w14,500 BP) (Mooser, 1967) to around 9000 BP (Krieger, 1964), to
as young as 7000 to 9000 BP (Armillas, 1964). Although the Basin of
Mexico is very rich in terms of mammoth nds, no Clovis points
have been found in the area.

Aveleyra de Anda and Maldonado-Koerdell, 1952, 1953). The skull


and tusks of an adult animal that had reached maximum growth
were overturned during butchery and some of the bones had deep
cut marks. The skeleton was incomplete and was lacking sections of
the mandible, both humeri, the right ulna, both radii, the left femur
and the right scapula. All the bones were displaced from their
correct anatomical position, apart from the leg bone. Of interest is
that the cranium was inverted, possibly for ease of brain extraction.
The skull of another mammoth excavated by Arellano (1946) close
to Tepexpan was in a similar position. The Iztapan mammoth was
found at depths from 1.80 to 2.25 m from the surface which was at
2238 m. Three lithics were found in direct association with the
mammoth bones (Fig. 11b).
5.3. New Santa Isabel Iztapan II stratigraphy and tephrochronology
The new trench stratigraphy is presented in Fig. 13, showing a
sequence of lacustrine sediments, with abundant ostracods, interbedded with several volcanic ash layers. The tephra geochemistry is
given in Table 3. The lowest tephra (Sample Santa 2) is 17.0 cm ne
sand black ash with a mean SiO2 of 56% (with a uniform, basalticandesitic composition). This is interpreted here as the GBA
(Mooser, 1967) or Tlahuac ash (Ortega and Newton, 1998). Between
260 and 270 cm there is a light grey sandy ash (Sample Santa 10)
that is chaotically mixed with lake sediments. This tephra has a
mean SiO2 of 67%, and currently its source is unknown. Its
geochemistry indicates it was neither the PWA nor UTP tephra
markers (Table 4). Within the green lake clay at around 220 cm in
the new trench section there is a layer with pockets of ash, pumice
and clay up to 2.2 cm across with a mixed tephra shard population
(samples Santa 12 and 13) with values of SiO2 of 58%, 64% and 71%.
Associated with this horizon are mud balls. The mixed tephra
population and the mud balls indicate reworking of tephra populations, including the PWA and lake sediments into the marginal,
nearshore lake environment.
There appears to be no in situ PWA in this trench as originally
described by Mooser and Gonzalez-Rul (1961; Fig. 12), but without

5.2. Previous work at Santa Isabel Iztapan Mammoths


Santa Isabel Iztapan I: In March 1952, following a chance nd
by workers opening a drainage ditch close to Santa Isabel Iztapan, a
mammoth skeleton was excavated by the Instituto Nacional de
Antropologia e Historia (INAH). It was completely embedded in
green clays (Aveleyra de Anda and Maldonado-Koerdell, 1952,
1953). The mammoth bones were disarticulated and appeared to
have been butchered in situ. Six lithic points (Fig. 11a) were found in
direct association with the skeleton. The other artefacts included a
scraper, a knife and a ne prismatic blade, all composed of obsidian
and an endscraper and a retouched int blade. One of the
mammoth femurs was a short distance from the remaining bones
as if pulled aside for butchering and approximately 80% of the total
bones were recovered.
Santa Isabel Iztapan II: In May 1954 the construction of another
ditch resulted in a second mammoth kill site, approximately
350 m south of Santa Isabel Iztapan I. Here, lithic points had been
used during mammoth butchering. In this case one of the posterior
legs of the animal was found in anatomical position, showing that it
had probably been trapped in the lake clay (Aveleyra de Anda, 1955;

Fig. 12. Tephra markers from the Santa Isabel Iztapan Mammoth II, Peon Woman III
and Tepexpan Man, modied after Mooser and Gonzalez-Rul (1961). PWA Pumice
with Andesite and UTP Upper Toluca Pumice. FH: Fossil human, F: Mammoth Find, S:
Soils, SA: Archaeological soils, T: Travertine, C: Caliche layer, AC: Consolidated sands, B:
Bentonite clays.

Please cite this article in press as: Gonzalez, S., et al., Paleoindian sites from the Basin of Mexico: Evidence from stratigraphy, tephrochronology
and dating, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.015

S. Gonzalez et al. / Quaternary International xxx (2014) 1e16

13

Fig. 13. Stratigraphy in new trench, Santa Isabel Iztapan Mammoth II, showing the position of samples taken and with M the position of the mammoth.

detailed tephra geochemistry and radiocarbon dating, which was


not available at the time, the most obvious feature in the layer
would have been the presence of pumices. The lack of in situ PWA
tephra has also been noted at the Tepexpan (Lamb et al. 2009) and
Tocuila sites (Gonzalez et al. 2014), in similar marginal lake locations. The youngest tephra (Sample Santa 16) is a 7 cm dark grey/
black basaltic ash with abundant ostracods, dominated by tephra
with a mean SiO2 composition of 58%, although there a few shards
of rhyolitic composition. We interpret this ash as probably equivalent to layer Toc-6 in Tocuila due to its composition and stratigraphic position, although we need more data to conrm this
conclusion. This layer was dated at Tocuila at w10,800  50 BP
(Gonzalez et al. 2014). A sample taken from a mammoth molar from
Santa Isabel Iztapan 2 (bone DP-412) could not be radiocarbon
dated because of low collagen content.
5.4. Signicance of the results
The Santa Isabel Iztapan I and II Mammoth sites are important
because of the undoubted association of unuted lithics and
mammoths. It was proposed previously that both were below the
PWA tephra marker with dates older than 14,500 BP, but our new
work has shown that there is no PWA tephra in situ in our new
trench and that in the stratigraphic layer (Samples Santa 12 and 13)
where there were pumice fragments, they are reworked. The layer
is a mixed tephra horizon with PWA and other ashes in a marginal
lake environment. However on top of the Santa Isabel Iztapan II
mammoth there is also a basaltic-andesitic tephra (sample Santa
16) that was dated at the Tocuila site at w10,800 BP. This means
that the age of both the mammoths and associated lithics are after
14,500 BP to 10,800 BP, before the onset of the Younger Dryas,
overlapping in time with the Clovis lithic tradition in the USA and
perhaps even older. This timing suggests that the Scottbluff, Lerma,
and Angostura lithics are likely to have originated rst in Central
Mexico, before spreading to the Great Plains of USA.
6. Discussion and conclusions
In the American Continent there are very few directly dated
Paleoindians;
examples
include
Buhl
Woman,
Idaho

10,675  95 BP; Spirit Cave Man, 9415 BP; Wizards Beach Man
9225 BP and Kennewick Man at 8410  60 BP. The oldest human in
the Basin of Mexico is the Peon Woman III skeleton dated to
10,755  75 BP but there are also human remains found in submerged caves in the Yucatan peninsula with comparable dates
between 9 and 11,000 BP (Gonzlez Gonzlez et al., 2006). The
Tlapacoya I cranium with a date of 10,200  65 BP is also important
for any discussion of humans in the New World. There are also
indirectly dated human skulls from the basin using tephrochronology: Chimalhuacan Man and Metro Man dated by association with the UTP tephra at w10,500 BP.
Our new stratigraphic work at Tlapacoya has produced no
further evidence to support early humans at 24,000 BP. However, it
is clear that the locality is an important Paleoindian site, because
one of the two human crania found previously has produced a
direct radiocarbon date of Late Pleistocene age. As both human
skulls are dolicocephalic, it means that there was a Paleoindian
population before 10,000 BP in Central Mexico with different cranial morphologies from Modern Amerindian populations. This
Paleoindian population includes the long skulls found associated
with the UTP ash at Chimalhuacan and El Metro sites.
We emphasize the importance of volcanic eruptions and volcanic hazards with early human presence in the Basin of Mexico.
We know that humans were suffering the consequences of major
Plinian volcanic eruptions during the Late Pleistocene. This major
environmental control and the implications for human and animal
populations living at the time has only been recently been incorporated into archaeological interpretations in Central Mexico during Paleoindian times (e.g. Gonzalez and Huddart, 2007).
The period between 15,000e10,000 BP in the Basin of Mexico
was characterised by extensive volcanic ash falls. There are two
main tephra markers, the PWA and the UTP that have been identied at the El Peon Woman III, Tlapacoya, Tocuila and Santa Isabel
Iztapan Mammoth II sites. However for the PWA ash there is also
evidence for multiple phases of deposition and considerable
thicknesses in the basin because there is extensive reworking and
redeposition of this ash on the hill slopes and into the margins of
the Texcoco and Chalco Lakes. Thus, great care is required when
interpreting volcanic ashes in marginal lake sites, where only by
using detailed geochemistry studies of single tephra grains,

Please cite this article in press as: Gonzalez, S., et al., Paleoindian sites from the Basin of Mexico: Evidence from stratigraphy, tephrochronology
and dating, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.015

14

S. Gonzalez et al. / Quaternary International xxx (2014) 1e16

Table 3
Summary of Santa Isabel Iztapan II Tephra Shard Geochemistry.
Sample

Mean SiO2%

S.D.

Variance

Range

Santa
Santa
Santa
Santa
Santa
Santa

17
16
4
4
3
11

56.61
67.87
58.04
63.99
72.60
57.92

0.74
1.83
0.98
1.37
2.54
1.86

0.54
3.33
0.96
1.87
6.47
3.45

54.17e57.35
64.2e70.91
56.92e58.95
62.36e65.71
69.75e74.63
53.44e60.28

2
10
13
13
16
16

detailed dating and good vertical and horizontal exposure of the


stratigraphic layers is it possible to identify those tephras which are
truly in situ, or reworked (Lamb et al. 2009; Gonzalez et al. 2014).
At Tlapacoya the evidence is unconvincing for a true lithic
assemblage. We postulate that the andesite akes recovered from
the original excavations are geofacts derived from the local bedrock
that fractures into naturally sharp akes. The small number of
obsidian akes from the original excavations was likely introduced
by rodent burrowing from above and the two worked bone
fragments seem to be simply broken bone. The proposed Pleistocene lake beach gravels and hearths from the original excavations
appear to us as simply local angular pebble-gravel, scree, and burnt

Table 4
Comparison of Geochemistry of Main Tephra Markers in the Basin of Mexico.
Santa Isabel Iztapan, Santa 10, n [ 16
SiO2 mean 67.87%, SD 1.83, variance 3.33, range 64.20 to 70.91%
FeO mean 2.45%, SD 1.47, variance 2.16, range 1.38 to 5.68%
MgO mean 0.53%, SD 0.16, variance 0.03, range 0.32 to 0.84%
CaO mean 2.64%, SD 0.81, variance 0.66, range 1.37 to 4.46%
Tocuila, Toc A1C, Tlahuac (GBA), n [ 18,
SiO2 mean 57.00%, SD 1.24, variance 1.54, range 54.62 to 59.33%
FeO mean 6.45%, SD 2.11, variance 4.46, range 1.14 to 8.72%
MgO mean 2.37%, SD 1.07, variance 1.14, range 0.2 to 5.36%
CaO mean 1.82%, SD 0.15, variance 0.02, range 1.51 to 1.95
Tocuila, Toc D2 (UTP) n [ 15
SiO2 mean 70.29%, SD 1.19, variance 1.41, range 68.76 to 73.26%
FeO mean 1.802%, SD 0.21, variance 0.45, range 1.22 to 2.09%
MgO mean 0.39%, SD 0.06, variance 0.004, range 0.18 to 0.45%
CaO mean 1.72%, SD 0.39, variance 0.153, range 0.74 to 2.46%
Tocuila, Toc D3 (UTP reworked) n [ 9,
SiO2 mean 70.96%, SD 2.13, variance 4.52, range 69.52 to 76.09%
FeO mean 1.62%, SD 0.034, variance 0.11, range 0.95 to 1.87%
MgO mean 0.35%, SD 0.95, variance 0.009, range 0.17 to 0.44%
CaO mean 1.59%, SD 0.53, variance 0.28, range 0.45 to 1.93%
Metro Man Skull (UTP) n [ 13
SiO2 mean 68.21%, SD 3.11, variance 9.67, range 63.06 to 71.86%
FeO mean 2.89%, SD 1.84, variance 3.39, range 0.3 to 5.35%
MgO mean 1.10%, SD 1.19, variance 1.43, range 0.01 to 2.09%
CaO mean 2.37%, SD 1.16, variance 1.35, range 0.21 to 4.42%
Chimalhuacan Man (UTP, reworked?) n [ 11,
SiO2 mean 68.69%, SD 4.59, variance 21.02, range 62.25 to 77.66%
FeO mean 3.19%, SD 1.82, variance 3.32, range 0.53 to 5.46%
MgO mean 0.95%, SD 1.36, variance 1.86, range 0.09 to 2.10%
CaO mean 2.85%, SD 1.36, variance 1.86, range 1.29 to 4.94%
Tlapacoya C9 (UTP in situ) n [ 7,
SiO2 mean 69.62%, SD 0.56, variance0.31, range 68.62 to 70.38%
FeO mean 1.76%, SD 0.01, variance 0.01, range 1.61 to 1.92%
MgO mean 0.41%, SD 0.03, variance 0.001, range 0.37 to 0.45%
CaO mean 1.82%, SD 0.15, variance 0.02, range 1.51 to 1.95%
Tlapacoya A2 (PWA), n [ 11,
SiO2 mean 59.07%, SD 2.17, variance 4.71, range 56.73 to 60.89%
FeO mean 6.97%, SD 1.13, variance 1.28, range 5.51 to 9.42%
MgO mean 3.12%, SD 1.39, variance 1.94, range 1.75 to 6.92%
CaO mean 5.33%, SD 0.83, variance 0.69, range 4.13 to 6.69%
Tephra 11 (UTP) from Chalco Lake
(Ortega-Guerrero and Newton, 1998) n [ 10,
SiO2 mean 71.2%, SD 1.27, variance 1.60, range 69.71 to 74.24%
FeO mean 1.86%, SD 0.13, variance 0.02, range 1.62 to 2.11%
MgO mean 0.39% SD 0.043, variance 0.002, range 0.32 to 0.43%
CaO mean 1.74%, SD 0.2, variance 0.04, range 1.46 to 2.04%

vegetation. The UTP layer was in situ in our Trench C. It is in this


layer that the stratied human cranium was reported to be
embedded. This layer has been dated to 9920  250 BP (Garcia
Barcena, 1986) and which agrees more or less with the AMS C14
date for the other unstratied human cranium at 10,200  65 BP
(Gonzalez et al. 2003).
The shallow marginal Lake Texcoco at the Tocuila Mammoths
site contains possible evidence for a thin Younger Dryas meteorite
airburst layer, mixed with volcanic ash, dated to approximately
10,800 BP. Three hundred years later (w10,500 BP) the large UTP
Plinian volcanic eruption also severely disrupted the basin
ecosystem. The large volume of tephra caused partial damming and
change in the drainage and the input of ood and lahar deposits
into the lakes. The timing of these later events seems synchronous
or post-dates the age of the UTP eruption occurring within the
Younger Dryas period. Into the lahars was incorporated a vertebrate
assemblage that included mainly mammoth (Mammuthus columbi)
skeletons and a few bone tools. The UTP ash has been found
associated with other Paleoindian skeletons in the Basin of Mexico
and it seems likely that both the potential meteorite airburst and
the UTP Plinian eruption caused widespread environmental
disruption of the ecosystem, the death of humans, and the death of
megafaunal populations, that were already weakened by human
predation and climate change during the Younger Dryas period.
There is little evidence for megafauna presence after this large
Plinian eruption in the Basin of Mexico and the possible meteorite
airburst. Afterwards, there seems to be a gap in the basin archaeological record, with a discontinuity in the hunter-gatherer, Late
Pleistocene Paleoindian populations. This hiatus in human occupation and changes in site use have been proposed for several sites
around the world at 10.8 ka (Wittke et al., 2013).
Evidence for re-occupation in the basin is not observed until
approximately 4500 BP. This is from the dating of the Preceramic
site of San Vicente Chicoloapan, showing a site with incipient
agriculture and food processing (metates) (Gonzalez et al. 2003).
There is an urgent need to establish if this archaeological gap is real
or apparent as perhaps we have been looking in the wrong lake
locations. The problem is that the Preceramic occupation sites on
the shores of the early Holocene lakes in the basin are likely to be
buried today at depths between 2 and 3 m under sediments, ashes,
and soils.
The new tephra dates given to the mammoths from Santa Isabel
Iztapan I and II (between 14,500 and 10,800 BP) indicates that the
fossil nds were probably butchered during the Clovis time period
but by a culture lacking Clovis points. This suggests that the Lerma
and Scottsbluff lithic types found in Central Mexico are likely to be
older than those from cultures in the USA, found with bison kill
sites.
Acknowledgements
We are very grateful to Don Celso Ramirez and family for access
to the Tocuila site and Museum during the last 15 years and Jos
Concepcin Jimnez Lpez (INAH, Mexico) for access and sampling
of the Preceramic Humans collection at the Museo Nacional de
Antropologa e Historia, Mxico. We acknowledge the help during
eldwork and discussions with Alan Turner and Graham Sherwood
from LJMU in the early development of this work. We thank the
NERC Electron Microprobe Unit at Edinburgh University, especially
Peter Hill, David Steel and Anthony Newton for access to the facilities and support over the years during the tephra analysis.
Financial help during eldwork was obtained from Liverpool John
Moores University and NERC, grant NE/C519446/1 and from the
Coordinacin de Investigacin Cientca de la Universidad
Michoacana de San Nicols de Hidalgo, Mexico.

Please cite this article in press as: Gonzalez, S., et al., Paleoindian sites from the Basin of Mexico: Evidence from stratigraphy, tephrochronology
and dating, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.015

S. Gonzalez et al. / Quaternary International xxx (2014) 1e16

References
Alvarez, T., 1986. Fauna pleistocnica. In: Lorenzo, J.L., Mirambell, L. (Eds.), Tlapacoya: 35000 Aos de Historia del lago de Chalco, Instituto de Antopologa e
Historia, Mexico DF, pp. 173e203.
Arce, J.L., Mcias, J.L., Vazquez Selem, L., 2003. The 10.5 ka Plinian eruption of
Nevado de Toluca, Mexico, stratigraphical and hazard implications. Geological
Society of America Bulletin 115, 230e248.
Arellano, A.R.V., 1946. El Elefante fosil de Tepexpan y el Hombre primitivo. Revista
Mexicana de Estudios Antropologicos 8, 89e94.
Armillas, P., 1964. Northern Mesoamerica. In: Jennings, J.D., Norbeck, E. (Eds.),
Prehistoric Man in the New World. University of Chicago Press, Chicago and
London, pp. 291e329.
Arroyo-Cabrales, J., Johnson, E., Morett, L., 2001. Mammoth bone technology in the
Basin of Mexico. In: Cavarratta, G., Gioia, P., Mussi, M., Palombo, M.R. (Eds.),
Proceedings of 1st International Congress The World of Elephants, Rome,
pp. 419e423.
Arroyo-Cabrales, J., Gonzalez, S., Morett, A.L., Polaco, O., Sherwood, G., Turner, A.,
2003. The Late Pleistocene paleoenvironments of the Basin of Mexico- evidence
from the Tocuila Mammoth site. Deinsea 9, 267e272.
Aveleyra Arroyo de Anda, L., 1955. El Segundo Mamut Fsil de Santa Isabel Iztapan,
Mxico y Artefactos Asociados. Direccin de Prehistoria, No.1. INAH, Mexico, 61 pp.
Aveleyra Arroyo de Anda, L., 1956. The second Mammoth and associated artifacts at
Santa Isabel Iztapan, Mexico. American Antiquity 22, 12e28.
Aveleyra de Anda, L., 1964. The primitive Hunters. In: Handbook of Middle American Indians V. I, pp. 348e412. Austin, Texas.
Aveleyra Arroyo de Anda, L., Maldonado-Koerdell, M., 1952. Asociacin de Artefactos con Mamut en el Pleistoceno Superior de la Cuenca de Mexico. Revista
Mexicana de Estudios Antropolgicos 13, 3e29.
Aveleyra Arroyo de Anda, L., Maldonado-Koerdell, M., 1953. Association of Artifacts
with Mammoth in the Valley of Mexico. American Antiquity 18, 332e340.
Brcenas, M., del Castillo, A., 1887. Noticia acerca del hallazgo de restos humanos
prehistricos en el Valle de Mxico. In: La Naturaleza, Mxico, 1 series, vol. 7,
pp. 257e264.
Bloomeld, K., Valastro, S., 1977. Late Quaternary Tephrachronology of Nevado de
Toluca Volcano, Central Mexico. Overseas Geology and Mineral Resources 46,
1e15.
Bradbury, J.P., 1971. Paleolimnology of Lake texcoco, evidence from diatoms.
Limnology and Oceanography 16, 180e200.
Bradbury, J.P., 1986. Paleolimnologa del lago de Chalco, Mxico. El medio ambiente
litoral. In: Lorenzo, J.L., Mirambell, L. (Eds.), Tlapacoya: 35000 Aos de Historia
del lago de Chalco. Instituto de Antroploga e Historia, Mexico DF, pp. 167e172.
Bradbury, J.P., 1989. Late Quaternary lacustrine paleoenvironments in the Cuenca de
Mxico. Quaternary Science Reviews 8, 75e100.
Caballero Miranda, M.E., 1997. The Last Glacial Maximum in the Basin of Mexico:
the diatom record between 34,000 and 15,000 years BP from Lake Chalco.
Quaternary International 43/44, 125e136.
Carballal-Staedtler, M. (Ed.), 1997. A propsito del Quaternario, Homenaje al Professor Francisco Gonzalez Rul, Direccin de Salvamento Arquelgico. INAH,
Mexico, DF.
Firestone, R.B., West, A., Kennett, J.P., Becker, L., Bunch, T.E., Revay, Z.S., Schultz, P.H.,
Belgya, T., Kennett, D.J., Erlandson, J.M., Dickenson, O.J., Goodyear, A.C.,
Harris, R.S., Howard, G.A., Kloosterman, J.B., Lechler, P., Mayewski, P.A.,
Montgomery, J., Poreda, R., Darrah, T., Que Hee, S.S., Stick, A., Topping, W.,
Wittle, J.H., Wolbach, W.S., 2007. Evidence for an extraterrestrial impact 12,900
years ago that contributed to the megafaunal extinctions and the Younger Dryas
cooling. Proceedings of the National Academy of Sciences, USA 104, 16016e
16021.
Garca-Brcena, J., 1986. Algunos aspectos cronolgicos. In: Lorenzo, J.L.,
Mirambell, L. (Eds.), Tlapacoya: 35000 Aos de Historia del lago de Chalco,
Coleccin Cientca, vol. 115. Instituto de Antropologa e Historia, Mexico DF,
pp. 219e224.
Gonzlez Gonzlez, A.H., Rojas Sandoval, C., Terrazas Mata, A., Benavente
Sanvicente, M., Stinnesbeck, W., 2006. Estudio preliminar de tres crneos
tempranos procedentes de cuevas sumergidas de las costa este Quintana Roo.
In: Jimnez Lpez, J.C., Polaco, O.J., Martnez Sosa, G., Hernndez Flores, R.
(Eds.), 2o Simposio Internacional el hombre temprano en Amrica. INAH,
Mxico, pp. 189e197.
Gonzalez, S., Huddart, D., Morett-Alatorre, L., Arroyo-Cabrales, J., Polaco, O.J., 2001.
Volcanism and Early Humans in the Basin of Mexico during the late Pleistocene/
Early Holocene. In: Cavarratta, G., Gioia, P., Mussi, M., Palombo, M.R. (Eds.),
Proceedings of 1st International Congress The World of Elephants, Rome,
pp. 704e706.
Gonzalez, S., Jimnez-Lpez, J.C., Hedges, R., Huddart, D., Ohman, J.C., Turner, A.,
2003. Earliest humans in the Americas: new evidence from Mexico. Journal of
Human Evolution 44, 379e387.
Gonzalez, S., Morett-Alatorre, L., Huddart, D., Arroyo-Cabrales, J., 2006. Mammoths
from the Basin of Mexico: stratigraphy and radiocarbon dating. In: Jimnez
Lpez, J.C., Pompa y Padilla, J.A., Gonzalez, S., Ortiz, F. (Eds.), El Hombre Temprano en America, Proceedings of the 1st International Symposium Early
Humans in America, Coleccon Cientica, Serie Antropologia, vol. 500. INAH,
Mexico City, pp. 263e274.
Gonzalez, S., Huddart, D., 2007. Paleoindians and megafaunal extinction in the Basin
of Mexico: the role of the 10.5 k Upper Toluca Pumice Eruption. In: Grattan, J.,

15

Torrance, R. (Eds.), Living under the Shadow: Cultural Impacts of Volcanic


Eruptions, One World Archaeology Series 53. Left Coast Press, Walnut Creek,
California, pp. 90e106.
Gonzalez, S., Huddart, D., 2008. The late pleistocene human occupation of Mexico.
In: Adauto, de Arajo et al, J.G. (Eds.), 11 Simposio Internacional. O Povamiento
das Amricas 2006. Sai Raimundo Nonanto, FUMDHAMentos, vol. V11, pp. 236e
259.
Gonzalez, S., Huddart, D., Israde-Alcntara, I., Dominguez-Vazquez, G., Bischoff, J.,
2014. Tocuila mammoths, Basin of Mexico: Late Pleistocene-Early Holocene
stratigraphy and the geological context of the bone assemblage. Quaternary
Science Reviews. http://dx.doi.org/10.1016/j.quascirev.2014.02.003.
Haynes, C.V., 1969. The earliest Americans. Science 166, 709e715.
Israde -Alcntara, I., Bischoff, J.L., Dominguez-Vasquez, G., Hong-Chun, L.,
DeCarli, P.S., Bunch, T.E., Wittke, J.H., Weaver, J.C., Firestone, R.B., West, A.,
Kennett, J.P., Mercer, C., Sinjing, X., Richman, E.K., Kinzie, C.R., Wolbach, W.S.,
2012. Evidence for Central Mexico supporting the Younger Dryas extraterrestrial impact hypothesis. Proceedings of the National Academy of Sciences, USA
109 (34), E738eE747.
Jimnez-Lpez, J.C., Hernndez Flores, R., Martnez Sosa, G., Saucedo Arteaga, G.,
2006. La Mujer del Peon III. In: Jimnez-Lpez, J.C., Gonzlez, S., Pompa y
Padilla, J.A., Ortiz-Pedraza, F. (Eds.), El Hombre Temprano en Amrica, Proceedings of the 1st International Symposium Early Humans in America, Serie
Antropologa Coleccin Cientca, vol. 500. INAH, Mexico City, Mexico, pp. 49e
66.
Johnson, E., Morrett, A.L., Arroyo-Cabrales, J., 2001. Late pleistocene bone technology at tocuila, Basin of Mexico. Current Research in the Pleistocene 18, 13e15.
Kennett, D.J., Kennett, J.P., West, A., Mercer, C., Que Hee, S.S., Bement, L., Bunch, T.E.,
Sellers, M., Wolbach, W.S., 2009. Nanodiamonds in the Younger Dryas boundary
sediment layer. Science 323, 94.
Krieger, A.D., 1964. Early Man in the New World. In: Jennings, J.D., Norbeck, E. (Eds.),
Prehistoric Man in the New World. University of Chicago Press, Chicago and
London, pp. 23e84.
Lamb, A.L., Gonzalez, S., Huddart, D., Metcalfe, S.E., Vane, C.H., Pike, A.W.G., 2009.
Tepexpan Palaeoindian site, Basin of Mexico: multi-proxy evidence for environmental change during the late Pleistoceneelate Holocene. Quaternary Science Reviews 28, 2000e2016.
Lozano-Garca, M., Ortega-Guerrrero, B., Caballero-Miranda, M., UrrutiaFucugauchi, J., 1993. Late pleistocene and holocene paleoenvironments of
chalco Lake, Central Mexico. Quaternary Research 40, 332e342.
Lorenzo, J.L., Mirambell, L. (Eds.), 1986a. Tlapacoya: 35000 Aos de Historia del lago
de Chalco, Coleccin Cientca 115. Instituto de Antropologa e Historia, Mexico
DF.
Lorenzo, J.L., Mirambell, L., 1986b. Mamutes excavados en la Cuenca de Mxico
(1952e1980), Cuaderno de Trabajo, vol. 32, 151 pp., Mxico.
Martinez del Rio, P., 1952. El Mamut de Santa Isabel Iztapan. Cuadernos Americanos
9, 149e170.
Mirambell, L., 1967. Excavaciones en un sitio pleistocnico de Tlapacoya, Mxico.
Boletin, Instituto Nacional de Antrolopoga y Hitsoria, Mxico 29, 37e41.
Mirambell, L., 1978. Tlapacoya: a late Pleistocene site in central Mexico. In:
Bryan, A.L. (Ed.), Early Man in America from a Circum-Pacic Perspective.
University of Alberta, Alberta, pp. 221e230.
Mirambell, L., 1986a. La excavaciones. In: Lorenzo, J.L., Mirambell, L. (Eds.), 1986
Tlapacoya: 35000 Aos de Historia del lago de Chalco, Mexico DF, Instituto de
Antropologa e Historia, Tlapacoya: 35000 Aos de Historia del lago de Chalco,
Mexico DF. Instituto de Antropologa e Historia, Mexico DF, pp. 13e54.
Mirambell, L., 1986b. Restos culturales en horizontes pleistocnicos. In: Lorenzo, J.L.,
Mirambell, L. (Eds.), 1986 Tlapacoya: 35000 Aos de Historia del lago de Chalco,
Coleccin Cientca, vol. 115. Instituto de Antropologa e Historia, Mexico DF,
pp. 207e217.
Mooser, F., 1967. Tefracronologia de la Cuenca de Mexico para los ultimos treinta mil
aos. Boletin del INAH de Mexico 30, 12e15.
Mooser, F., 1997. Nueva fecha para la tefracronologia de la Cuenca de Mxico. In:
Carballal-Staedtler, M. (Ed.), A propsito del Quaternario. Direccin de Salvamento Arqueolgico. INAH, Mexico, pp. 137e141.
Mooser, F., Gonzlez-Rul, F., 1961. Erupciones volcnicas y el hombre primitivo en la
Cuenca de Mxico. In: Homenaje a Pablo Martinez del Rio en el XXV Aniversario
de la edicin de Los Orgenes Americanos. INAH, Mexico, pp. 137e141.
Morett, L., Arroyo-Cabrales, J., Polaco, O.J., 1998. Tocuila Mammoth site. Current
Research in the Pleistocene 15, 118e120.
Newberry, J., 1887. Discusiones acerca del Hombre de Peon. In: La Naturaleza, Ia
Series, vol. 7, pp. 284e285. Mxico.
Ortega-Guerrero, B., Newton, A.J., 1998. Geochemical characterization of late
pleistocene and holocene tephra layers from the Basin of Mexico, Central
Mexico. Quaternary Research 50, 90e106.
Pichardo, M., 2000. Redating Iztapan and Valsequillo, Mexico. Radiocarbon 42,
305e310.
Pompa y Padilla, J.A., 1988. Nueva evidencia en Mexico: datos preliminares del
hombre de Chimalhuacan. In: Gonzalez Jacome, A. (Ed.), Origenes del hombre
americano, SEP. Cien de Mexico, pp. 177e207.
Reyes, A.E., 1923. Los Elephantes de la Cuenca de Mxico. Revista Mexicana de
Biologia 111 (6), 227e244.
Roberts, C., Manchester, K., 1997. The Archaeology of Disease, second ed. Cornell,
New York.
Romano, A., 1964. Restos humanos precermicos de Mxico. Antropologa Fsica,
Epoca Prehispnica. In: Mxico: Panorama Histrico y Cultural, Mxico.

Please cite this article in press as: Gonzalez, S., et al., Paleoindian sites from the Basin of Mexico: Evidence from stratigraphy, tephrochronology
and dating, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.015

16

S. Gonzalez et al. / Quaternary International xxx (2014) 1e16

Romano, A., 1970. Preceramic human remains. In: Handbook of Middle American
Indians, Austin, Physical Anthropology. University of Texas Press, 9.
Romano, A., 1974. Restos seos humanos precermicos de Mxico. In: Romero
Molina, J. (Ed.), Mxico: Panorama Histrico y Cultural, volume III. Instituto de
Antropologa e Historia, Mexico City, pp. 29e81.
Siebe, C., Schaaf, P., Urrutia-Fucugauchi, J., 1999. Mammoth bones embedded in a
late Pleistocene lahar from Popocatpetl volcano, near Tocuila, central Mexico.
Bulletin of the Geological Society of America 111, 1550e1562.
Tankersley, K., 2002. In Search of Ice Age Americans. Gibbs Smith, Layton, Utah.
Urrutia-Fucugauchi, J., Martin del Pozzo, A.L., 1993. Implicaciones de los datos
paleomagneticos sobre la edad de la Sierra de Chichinautzin, cuenca de Mexico.
Geofsica Internacional 32, 523e533.
Urrutia-Fucugauchi, J., Lozano-Garca, M.S., Ortega-Guerrero, B., CaballeroMiranda, M., Hansen, R., Negendank, J.F.W., 1995. Palaeomagnetic and

palaeoenvironmental studies in the southern Basin of Mxico- 11 Late PleistoceneeHolocene Chalco lacustrine record. Geosica International 43, 33e53.
Wittke, J.H., Weaver, J.C., Bunch, T.E., Kennett, J., Kennett, D., Moore, A., Hillman, G.,
Tankersley, K., Goodyear, A., Moore, C., Daniel Jr., I.R., Ray, J., Lopinot, N.,
Ferraro, D., Israde-Alcntara, I., Bischoff, J., DeCarli, P., Hermes, R.,
Kloosterman, J., Revay, Z., Howard, G., Kimbel, D., Kletetschka, G., Nabelek, L.,
Lipo, C., Sakai, S., West, A., Firestone, R., June 2013. Evidence for Deposition of 10
Million Tonnes of Impact Spherules Across Four Continents 12,800 y Ago. www.
pnas.org/cgi/doi/10.1073/pnas.1301760110.
Wormington, H.M., 1957. Ancient Man in North America, Denver Museum of Natural History Popular Series No.4, fourth ed. Peerless Publishing Company,
Denver.

Please cite this article in press as: Gonzalez, S., et al., Paleoindian sites from the Basin of Mexico: Evidence from stratigraphy, tephrochronology
and dating, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.015

Quaternary International xxx (2014) 1e8

Contents lists available at ScienceDirect

Quaternary International
journal homepage: www.elsevier.com/locate/quaint

The role of plants in the early human settlement of Northwest South


America
s Loaiza a, b
Francisco Javier Aceituno a, *, Nicola
a
b

Grupo Medioambiente y Sociedad, Departamento de Antropologa, Universidad de Antioquia, Calle 67 No 53-108, AA 1226 Medellin, Antioquia, Colombia
Temple University Department of Anthropology, Philadelphia, PA 19119, USA

a r t i c l e i n f o

a b s t r a c t

Article history:
Available online xxx

This paper presents a synthesis on the existing of the role of plants on the adaptive strategies of human
groups that settled in Northwest South America since the Pleistocene/Holocene transition. To contextualize the analysis, a brief description of the Colombian Pleistocene sites is presented. The paper presents a broad description of the lithic technology, archaeobotanical record and radiocarbon dates. Plant
resources played a key role in the settling of human groups in the forests of the Neotropics. Furthermore,
it is suggested that for some areas there is evidence of cultivation as a strategy to increase the carrying
capacity of the surrounding environment.
2014 Elsevier Ltd and INQUA.

Keywords:
Colombia
Northwest South America
Early peopling
Use of plants
Tropical forests
Lithic technology

1. Introduction
Due to its geographical position, Northwest South America,
mostly corresponding to the current Colombian territory, is a
crucial area to address the issues of adaptive strategies, and the
process of early human dispersion in northern South America since
the Pleistocene/Holocene transition. The earliest date was found at
Pubenza (Fig. 1), an open air site located in the lowlands of the
Magdalena River Basin, where mastodon bones have been found
associated with eight stone akes in a layer dated at 16,460 420
BP (Van der Hammen and Correal, 2001). El Abra, a rock shelter
located 2600 m asl at the Sabana de Bogota (Eastern Cordillera)
(Fig. 1) where archaeologists recovered Holocene faunal remains
associated with unifacial lithic tools dated 12,400 160 BP, is the
second oldest site in Colombia (Correal et al., 1966e1969; Hurt
et al., 1977; Correal, 1986).
The Sabana de Bogota (Fig. 1) has one of the longest and most
complete occupation sequences in Northwest South America,
starting at ca. 12,400 BP through the XVI Century AD (Correal and
van der Hammen, 1977; Correal, 1981, 1986). Regarding the
earliest humans, Correal, the archaeologist who lead the research at
the Sabana de Bogota, suggested that between ca. 11,000 and
10,000 BP, semi-specialized hunters had adapted to the semi-open
environment of the Bogota highland (Correal, 1986, p. 119).

* Corresponding author.
E-mail
addresses:
aceitunob@hotmail.com,
(F.J. Aceituno), nloaiza@temple.edu (N. Loaiza).

csfjace@antares.udea.edu.co

The high frequency of faunal remains in the sites from Sabana de


Bogota including the Pleistocene megafauna remains (Cuvieronius
hyodon, Haplomastodon sp. and Equus amerhippus sp.) from Tibito
(11,740 110 BP) indicates that hunting was an important activity
that prevailed through the middle Holocene (Correal, 1982, 1986, p.
 pez
124e125). The emphasis on hunting was also suggested by Lo
(1999) writing about the middle Magdalena River Basin (Middle
Magdalena) (Fig. 1), where he stated that, so far, the lithic assemblages recovered suggest the existence of a tradition of
specialized hunters that reminds us of the denitions of Paleopez, 1999, p. 100). Along with the evidence
indian cultures (Lo
found at Sabana de Bogota and Middle Magdalena, the supercial
 pez, 1995, p.
recoveries of projectile points through Colombia (Lo
75), forged the widespread idea for Colombian archaeology that the
early human groups focused their economic strategies mainly on
hunting. During the 1960s, Reichel-Domatoff (1997 [1965] p. 40)
argued that Northwest South America environmental conditions
must have determined a strong emphasis on minor species hunting
strategies as well as in gathering during the Paleoindian stage. This
idea challenged the traditional view of Paleoindian human groups
as mainly megafauna hunters and was based on the fact that
neither the amount of projectile points nor the megafauna remains
associated with human activities in Colombia supported that idea.
By the Pleistocene/Holocene transition (ca.10,500 BP), archaeological data suggest that the adaptive strategies had strong
emphasis on plant resources. The aim of this paper is to present a
compilation of archaeobotanical and paleoecological data suggesting that, along with strategies such as hunting and perhaps
shing, plant resources were an important part of the adaptive

http://dx.doi.org/10.1016/j.quaint.2014.06.027
1040-6182/ 2014 Elsevier Ltd and INQUA.

Please cite this article in press as: Aceituno, F.J., Loaiza, N., The role of plants in the early human settlement of Northwest South America,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.06.027

F.J. Aceituno, N. Loaiza / Quaternary International xxx (2014) 1e8

Fig. 1. Colombian geography and archaeological areas distribution.

strategies that enabled human settlement and, possibly, the


incipient construction of cultural traditions that spread through
Colombia at the Pleistocene/Holocene transition (ca.10,500 BP)
until the early Holocene (ca. 8000 BP) (Aceituno et al., 2013).
2. The role of tropical forests in the early peopling of Central
America and Northwest South America
The Paleoindian evidence found in moist forests in Central and
South America allowed a debate around the role that this environments had in the peopling of the Neotropics during the nal Pleistocene, a time in which environments were different from today's,
mainly due to changes in temperature and rainfall (Piperno and
Pearsall, 1998; Ranere, 2006, 2008). Broadly speaking, the paleoecological data from Central America suggest the existence of
several types of landscapes: montane forest, thorn scrub, savannah,
tropical moist forests, etc. (Snarkis, 1979; Bush and Colinvaux, 1990;
Piperno and Pearsall, 1998; Ranere, 2008). This situation has been
studied in Panama, where the relation between paleoenvironments
and archaeological sites has been analyzed in order to suggest
subsistence strategies in tropical ecosystems for the earliest settlers
of the isthmus (Piperno and Pearsall, 1998; Ranere, 2006). Facing the
assumption that tropical forests must have posed a barrier for human migrations, Ranere (2006) suggested that human groups must
have taken advantage a wide range of ecosystems, which included
tropical forests. Ranere noted that hunting was the main subsistence

strategy at the beginning. With the decrease of game and demographic increase towards the nal Pleistocene, human strategies
shifted towards the exploitation of plant resources. The oldest indirect evidence for environmental management comes from a
sediment layer extracted from a core at La Yeguada Lake and dated
between ca 11,000 and 10,000 BP, where charred phytoliths of
grasses and Heliconia were recovered along with an abrupt increase
of the frequency of microscopic charcoal, suggesting forest burning
by human groups that took advantage of the natural resources of
this environment (Piperno, 1995; Piperno and Pearsall, 1998).
For Colombia, the most complete sequence of occupations
comes from the highlands of the Sabana de Bogota (about 500 km
SE of the PanamaeColombia border on the Pacic Ocean, ~2600 m
asl) in an Andean forest in today's environmental classication. The
environmental reconstructions of the Guantiva Interstadial, dated
ca 12,500 to 11,000 BP, indicate that the Andean moist forest that
was replaced by sub-Paramo vegetation during the Abra Interstadial (Correal, 1986; Marchant et al., 2002). The archaeological evidence for this time frame comes from El Abra II, Tequendama I and
Tibito (Table 1) (Correal and van der Hammen, 1977; Correal, 1981,
1986). Faunal remains as well as the associated lithic technology
suggest that hunting was the main economic strategy in the Sabana
de Bogota. There are no direct indicators of edible plant usage.
Nonetheless, Correal (1986, p. 121) mentions the presence of the
pollen record of the Dodoneae plant family towards ca. 10,000 BP
and associates it with forest clearance.

Please cite this article in press as: Aceituno, F.J., Loaiza, N., The role of plants in the early human settlement of Northwest South America,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.06.027

F.J. Aceituno, N. Loaiza / Quaternary International xxx (2014) 1e8


Table 1
14
C dates from Northwest South America at the Pleistocene/Holocene transition.
Site

Region

14

1s

Calib BC/ADa

Pubenza
n
El Jorda
El Abra II

Tibito
Tequendama I
El Abra II
Tequendama I
El Abra II
Tequendama I
Tequendama I
Torre 46 (Nare)
Torre 46 (Nare)
La Palestina 2
Torre 46 (Nare)
San Juan de Bedout
La Palestina 2
La Palestina 2
PIII0I-52
La Palestina 2
Tequendama I
Tequendama I
Tequendama I
El Guatn
El Jazmin
Sueva I
La Morena
San Isidro
San Isidro
Tequendama I
Tequendama I
La Palestina 1
n
El Jorda
Tequendama I
66PER001
La Morena
Salento 24
Sauzalito
Sauzalito
La Trinidad I
San Isidro
La Selva
Gachal
a
El Abra II
La Trinidad II
El Abra II
La Pochola
Sauzalito
~ a Roja
Pen
nova
Ge
~ ita
La Montan
~ a Roja
Pen
~ a Roja
Pen
Sitio 045
El Abra II
El Abra II
El Jazmn
Sitio 021
El Abra II
El Abra II
El Recreo
Galindo I
Nuevo Sol
La Selva
39 El Recreo Cancha
~ a Roja
Pen
39 El Recreo Cancha
~ ones de Bogota

Pen
Salento 21
El Antojo
Neusa
PIIIOP-59
Checua
La Chillona

Ro Magdalena
Cordillera Central
Sabana de Bogota
Sabana de Bogota
Sabana de Bogota
Sabana de Bogota
Sabana de Bogota
Sabana de Bogota
Sabana de Bogota
Sabana de Bogota
Magdalena medio
Magdalena medio
Magdalena medio
Magdalena medio
Magdalena medio
Magdalena medio
Magdalena medio
Porce medio
Magdalena medio
Sabana de Bogota
Sabana de Bogota
Sabana de Bogota
Cauca medio
Cauca medio
Sabana de Bogota
Ro Medelln
n
Altiplano Popaya
n
Altiplano Popaya
Sabana de Bogota
Sabana de Bogota
Magdalena medio
Central Cordillera
Sabana de Bogota
Cauca medio
Ro Medelln
Cauca medio
Ro Calima
Ro Calima
Cauca medio
n
Altiplano Popaya
Cauca Medio
Sabana de Bogota
Sabana de Bogota
Cauca medio
Sabana de Bogota
Cauca Medio
Ro Calima

Ro Caqueta
Cauca medio
Cauca medio

Ro Caqueta

Ro Caqueta
Porce medio
Sabana de Bogota
Sabana de Bogota
Cauca medio
Porce medio
Sabana de Bogota
Sabana de Bogota
Ro Calima
Sabana de Bogota
Cauca medio
Cauca medio
Cauca medio

Ro Caqueta
Cauca medio
Magdalena medio
Cauca medio
Cauca medio
Sabana de Bogota
Porce medio
Sabana de Bogota
Cauca medio

16,400
12,910
12,400
11,740
10,920
11,210
10,730
10,720
10,590
10,460
10,400
10,400
10,400
10,350
10,350
10,300
10,260
10,260
10,230
10,150
10,140
10,130
10,130
10,120
10,060
10,060
10,050
10,030
10,025
9990
9820
9760
9740
9730
9680
9680
9670
9600
9542
9530
9490
9360
9340
9333
9325
9312
9300
9250
9230
9230
9160
9125
9120
9050
9025
9020
8990
8810
8760
8750
8740
8740
8680
8550
8510
8480
8480
8430
8380
8370
8340
8200
8200

420
60
160
110
250
90
105
400
90
130
40
60
90
60
90
70
70
50
90
150
100
150
50
70
90
60
100
60
95
100
115
160
135
100
60
100
100
100
50
100
110
45
40
65
100
55
100
140
40
50
90
250
90
470
90
60
80
430
350
160
60
50
60
60
110
40
40
100
90
90
40
110
40

18,830
13,724
13,236
11,826
11,358
11,311
10,858
11,370
10,771
10,744
10,473
10,488
10,613
10,472
10,581
10,448
10,435
10,225
10,293
10,293
10,174
10,289
10,074
10,078
10,020
9881
10,027
9825
10,007
9880
9698
9768
9467
9371
9275
9296
9291
9258
9147
9220
9221
8755
8730
8759
8832
8719
8780
8849
8561
8572
8612
8879
8616
9554
8476
8322
8380
9177
8818
8253
7972
7952
7871
7683
7826
7586
7586
7603
7237
7580
7520
7521
7328

C date

References
16,869
13,255
12,121
11,312
10,193
10,889
10,564
9371
10,427
10,007
10,125
10,095
10,025
10,016
9980
9872
9806
9818
9670
9317
9372
9299
9526
9447
9322
9366
9310
9321
9295
9272
8837
8713
8747
8797
8837
8784
8782
8720
8750
8632
8556
8538
8532
8421
8292
8420
8295
8211
8312
8302
8243
7609
8202
7056
7938
7970
7935
7001
7027
7548
7597
7607
7586
7497
7292
7497
7497
7186
7187
7184
7313
6913
7078

Van der Hammen and Correal, 2001


Salgado, 1998
Hurt et al., 1977
Correal, 1981
Correal and van der Hammen, 1977
Hurt et al., 1977
Correal and van der Hammen, 1977
Correal and van der Hammen, 1977
Correal and van der Hammen, 1977
Correal and van der Hammen, 1977
 pez, 2008
Lo
 pez, 2008
Lo
 pez, 2008
Lo
 pez, 2008
Lo
 pez, 1989
Lo
 pez, 2008
Lo
 pez, 2008
Lo
Otero et al., 2006
 pez, 2008
Lo
Correal and van der Hammen, 1977
Correal and van der Hammen, 1977
Correal and van der Hammen, 1977
Restrepo, 2013
Aceituno and Loaiza, 2007
Correal, 1979
Santos, 2010
Gnecco, 2000
Gnecco, 2000
Correal and van der Hammen, 1977
Correal and van der Hammen, 1977
pez, 2008
CAIN-OCENSA, 1997 in Lo
Salgado, 1998
Correal and van der Hammen, 1977
Cano, 2004
Santos, 2010
Tabares and Rojas, 2000
Bray et al., 1988
Bray et al., 1988
Restrepo, 2013
Gnecco, 2000
Rodrguez, 2002
Correal, 1979
Hurt et al., 1977
Restrepo, 2013
Hurt et al., 1977
Aceituno refer per.
Bray et al., 1988
Cavelier et al., 1995; Gnecco, 2000
Restrepo, 2013
Restrepo, 2013
Cavelier et al., 1995; Gnecco, 2000
Mora, 2003
Castillo and Aceituno, 2006
Hurt et al., 1977
Hurt et al., 1977
Integral, 1997
Castillo and Aceituno, 2006
Hurt et al., 1977
Hurt et al., 1977
Herrera et al., 1992
Pinto, 2003
Restrepo, 2013
Aceituno and Loaiza, 2007
Herrera et al., 2011
Llanos, 1997
Herrera et al., 2011
 pez, 2008
Lo
Tabares and Rojas, 2000
Integral, 1997
Rivera, 1991
Cardona et al., 2007
Groot, 1992
Restrepo, 2013
(continued on next page)

Please cite this article in press as: Aceituno, F.J., Loaiza, N., The role of plants in the early human settlement of Northwest South America,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.06.027

F.J. Aceituno, N. Loaiza / Quaternary International xxx (2014) 1e8

Table 1 (continued )
Site

Region

14

1s

Calib BC/ADa

San Germ
an II
La Pochola
~ a Roja
Pen
39 El Recreo Cancha

Cauca medio
Cauca medio

Ro Caqueta
Cauca medio

8136
8095
8090
8030

65
55
60
80

7348
7299
7301
7177

C date

References
6628
7587
6823
6686

Aceituno and Loaiza, 2007


Aceituno and Loaiza, 2007
Mora, 2003
Herrera et al., 2011 a

All calibrated results have 2 sigma calibration with program Calib Rev 7.0.0 (data set used: intCal3.14c).

The amount of data to evaluate the relation between early


peopling and tropical environments in Colombia is still scarce, and
most of the earliest dates are closer to 10,000 BP. The ecological
conditions at Middle Magdalena (ca. 150 m asl) have not been well
established, but it is assumed that the areas where archaeological
sites older that ca.10,000 BP are located were covered by lowland
pez, 1999). For this region, the archaeologtropical rain forests (Lo
ical record is composed exclusively of lithic technology that has
been associated with hunting activities, probably inuenced by the
fact that it is the Colombian region where the highest amount of
 pez, 1995, 1999). Noneprojectile points has been recovered (Lo
theless, the forest coverage that is assumed may suggest that there
should have been some form of exploitation of plant resources that
has not been clearly identied.
At the Middle Cauca (Fig. 1), in the Central Cordillera (~1600 m
asl), pollen cores recovered suggest that by 10,120 BP (the time of
the earliest archaeological evidence), human groups encountered
an evergreen montane forest, resembling today's classication, but
with a larger presence of cold weather plant taxa (Aceituno and
Loaiza, 2007; Mercado, 2011). Further south along the Central
Cordillera, at the Popayan Plateau (Fig. 1) (~1650 m asl), environmental reconstruction using pollen cores extracted from the San
Isidro site suggest that the plant coverage was evergreen montane
forests by ca.10,300 BP when the site was rst occupied.
The Sabana de Bogota, Middle Magdalena, Middle Cauca, and
Popayan Plateau data are very important because they show the
important role that forests played in human dispersions in the
Colombian Andes. Pollen records as well as plant residues extracted
from lithic tools are the best evidence we have on early plant usage
for the Pleistocene/Holocene transition, a time characterized by
climatic and ecosystem changes that involved the reduction of
open areas and the expansion of tropical forests (Piperno and
Pearsall, 1998; Marchant et al., 2002).
3. Plant resources and human settling during the early
Holocene
By the beginning of the Holocene (ca. 10,000 BP) there is an
increase in the frequency of archaeological sites in the Colombian
~ a Roja
Andes. The only non-Andean site found of that age is Pen
(350 m asl) (Fig. 1) located in the middle Caqueta River Basin (a
tributary of the Amazon River) (Cavelier et al., 1995). Broadly
speaking, this distribution has been interpreted by Aceituno et al.
(2013) as an indicator of population expansion along the river
valleys that cross the Cordilleras of the northern Andes, as humans
were adjusting to the newer environmental conditions.
The archaeological record of the early Holocene in Northwest
South America is characterized by a) the spread distribution of sites
through the Cordilleras, b) the occupational redundancy (reoccupation) of sites, c) a lithic technology associated with broad spectrum economies, and d) the presence of microbotanical remains on
stone tools.
In southeast Colombia at the Calima river basin (Fig. 1) in the
Western Cordillera (~1750 m asl), two sites were found in evergreen montane forests, Sauzalito and El Recreo, dated between ca.

9700 and 8800 BP (Table 1) (Salgado, 1988e1990). In both sites,


archaeologists recovered plant processing tools such as unifacial
akes, grinding bases, handstones, and hoes that may have been
hafted (Herrera et al., 1988; Salgado, 1988e1990). This tool kit was
associated with several activities such as soil removal for cultivation, tuber exploitation, and extraction of starchy hearths of palms,
and was important in challenging the ruling Paleoindian model
(Cardale et al., 1989; Gnecco and Salgado, 1989). The recovery of
charred seeds of palms and Persea sp., along with palm, bamboo,
and maranta phytoliths supported the idea that plants played a key
role on the human economic strategies in this region (Piperno,
1985; Piperno and Pearsall, 1998). On a larger scale, Calima data,
combined the data from Panama with lithic technology associated
to the Tropical Forest Archaic (Ranere, 1980, p. 35), strengthens the
idea that human activities started shifting towards plant processing
at the beginning of the Holocene.
South of Calima, at the Popayan Plateau (Central Cordillera) lies
the site of San Isidro, (Fig. 1) in evergreen montane forest, and dated
between 10,050 100 and 9530 100 BP (Table 1) (Gnecco, 2000,
2003). The lithic material recovered at San Isidro consists of thousands of chert artifacts and some obsidian, composed of unretouched and retouched akes, lanceolate bifaces and preforms, as
well as plant processing tools (Gnecco, 1994, 2000, 2003; Gnecco
and Mora, 1997; Mora and Gnecco, 2003). This last class of tools
includes edge grounded cobbles (9), at mills (5), and a polished
axe (Gnecco, 2000). In addition to the artifacts, thousands of charred seeds were found that included Persea (cf. americana), Erythrina
(cf. edulis), Caryocar, Virola, several kinds of palms including Acrocomia (Piperno and Pearsall, 1998; Gnecco, 2000), as well as rinds of
Lagenaria sp. (Gnecco, 2003; Gnecco and Aceituno, 2006). From an
edge ground cobble, starch grains from cf. Xanthosoma/Ipomoea
and/or Manihot, and Maranta (cf. arundinacea), as well as grasses
and legumes, were identied (Piperno and Pearsall, 1998).
According to Gnecco (2000, 2003) people at San Isidro were
doing articial selection on preferable resources, cultivating
selected species of plants. Pollen of colonizer species such as
Plantago and grasses were interpreted as evidence of ecosystem
alteration in order to create open areas to concentrate resources,
following Posey's idea of resource islands (Gnecco, 2003; Gnecco
and Aceituno, 2006, p. 93). Plants concentration was a strategy
that aimed to increase the carrying capacity of tropical ecosystems,
anticipating the establishment of agriculture (Gnecco, 2000, 2003).
Gnecco (2000) inferred that San Isidro inhabitants were practicing,
perhaps since the nal Pleistocene, a behavior he described as
agrilocality: a strategy that entails selective plant manipulation and
cultivation.
Middle Cauca is located to the north (Fig. 1). It lies in the Middle
Basin of the Cauca River that separates the Central and Western
Cordillera. Over twenty sites dating from the Pleistocene/Holocene
transition to the middle Holocene have been found in the Western
Cordillera towards the central part of Middle Cauca (Table 1)
(Tabares and Rojas, 2000; Rodrguez, 2002; Cano, 2004; Tabares,
2004; Aceituno and Loaiza, 2007; Cano, 2008), seven of which
have occupations dating between ca. 10,000 and 8000 BP. All seven
sites are located between 1400 and 1600 m asl in a wet premontane

Please cite this article in press as: Aceituno, F.J., Loaiza, N., The role of plants in the early human settlement of Northwest South America,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.06.027

F.J. Aceituno, N. Loaiza / Quaternary International xxx (2014) 1e8

Fig. 3. A- Handstones: 1) El Jazmin Block 3 Level 15 (dated between 7080 50 and


5625 50 BP); 2) La Pochola Block 1 Level 13 code 733 (dated between 6903 45 and
6743 45 BP. Starch grains extracted see Table 2); 3) San German Block 1 Level 9
(dated 8136 65). B- Milling Stones: 1) La Pochola Block 1 Level 12 (dated 6743 45);
2) El Jazmin Block 3 Level 16 (dated between 7080 50 and 5625 50 BP).

For Middle Cauca, there is also an important record of starch


grains recovered from stone tools, supporting the importance of
plants in the settlement of this region. For the time frame between
ca. 10,000 and 8000 BP starch grains have been analyzed for three
sites, El Jazmin, La Pochola, and La Selva. Even though starch grain
analysis is still in progress we have preliminarily identied Dioscorea sp. (Fig. 4a) associated with a date of 8712 60 BP (La Selva),
Phaseoulus grains (Fig. 4b) (not the domesticated species) associated with dates of 8095 55 BP (La Pochola) and 8712 60 BP (La
Selva), and Calathea sp. associated with a date of 8660 55 BP (El
Jazmn) (Dickau, 2012 personal communication).
North of the Middle Cauca sites along the Central Cordillera lies
the Porce River Basin (Fig. 1), where ve sites dated between 10,260
BP and 8000 BP are located. From south to north, La Morena
(~2050 m asl) (Table 1) is located in an evergreen Andean area
(Santos, 2010) in the Medellin River Valley (Fig. 1), named Porce
river as it travels further north. The remaining four sites (Site 52,
041, 021, and 59) (Table 1) are located on a tropical rainforest below
1000 m asl (Castillo and Aceituno, 2006; Cardona, 2012; Otero and
Santos, 2012). Site 52 is dated 10,260 50 BP, and has yielded three

Fig. 2. 1. Hoes: 1) El Jazmn Block 1 Level 9 (code 433); 2) 021 (code 021-444); 3) La
Pochola Block 1 square B3 Level 16 (code 282); 4) El Jazmn (supercial recollection);
5) 059 Block 1, square 2A Level 6 (code 34 308); 6) 021 (code 021-181); 7) 045 (code
45-267); 7) 059 Block 1 square C1 Level 7 (code 8-496).

forest life zone (Espinal, 1990, p. 65). In broad terms, the lithic
technology basically consists of simple akes, axes/hoes (Fig. 2: 1, 3,
4) hand stones, and milling bases manufactured on local volcanic
rocks (Fig. 3: A, B). In addition to this, the lithic assemblage found at
El Antojo site is composed by thousands of quartz akes as well as a
preform. Recently, two stemmed projectile points have been
recovered in excavations, and one of them has been dated between
ca. 8000 BP and 8500 BP (Herrera et al., 2011).

Fig. 4. Starch grain. a) Dioscorea spp. La Selva (code 90); b) Phaseolus spp. La Pochola
(code 762).

Please cite this article in press as: Aceituno, F.J., Loaiza, N., The role of plants in the early human settlement of Northwest South America,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.06.027

F.J. Aceituno, N. Loaiza / Quaternary International xxx (2014) 1e8

chert tools (Otero and Santos, 2012, p. 60e62) that have been
associated to the Middle Magdalena based on the raw materials and
the technology (Aceituno et al., 2013). Also associated with the
Middle Magdalena, two projectile points manufactured on chert
were recovered from surface collections in the Medellin River
Valley (Aceituno et al., 2013). The other sites display technology
associated with Calima and Middle Cauca, described in detail
elsewhere (Aceituno, 2001) (Fig. 2: 2,5,6,7,8).
The main archaeobotanical evidence presented for this region is
the pollen record, and it is not clear in terms of directly indicating
plant use. Palm pollen has been identied at 021 and La Morena. At
La Morena, Phaseolus sp. pollen was identied (Santos, 2010).
Starch grains identied as Disocorea sp. dated between ca.10,000
and 9500 BP were also recovered (Santos, 2010). Most of the data
from the Porce River Basin date only to the middle Holocene, but
nonetheless they support the argument that plants played a very
important role in the human settlement of the northern Andes.
Outside the Andes in the middle Caqueta River Basin (Amazon
~ a Roja (Fig. 1), dated between 9250
Basin) is the open air site Pen
and 8100 BP (Table 1) (Cavelier et al., 1995; Gnecco and Mora, 1997;
Mora, 2003, p. 102; Mora and Gnecco, 2003). The lithic assemblage
in the preceramic levels was composed of unifacial akes, choppers, drills, handstones, milling stones, hammers, and anvils manufactured on local materials such as chert, quartz, and igneous
rocks (Cavelier et al., 1995, p. 31e32). Thousands of charred seeds
and other macrobotanical remains belonging to different genera of
palm trees, as well as wild fruit remains identied as Anaueria
brasiliensis, Parkia multijuga, Inga spp., Passiora quadrangularis,
and Caryocar spp. were also recovered at this site (Morcote et al.,
1998). It has been suggested that the large amount of charred
palm seeds suggest that this kind of plants were likely the subject of
some form of selective management (Cavelier et al., 1995, p. 36e41;
Morcote et al., 1998). Furthermore, the identication of Cucurbita
sp., Lagenaria siceraria, and Calathea sp. through phytoliths suggest
that these foreign plants were carried to this evergreen wet environment for cultivation (Piperno and Pearsall, 1998, p. 204e205).
4. Discussion
The archaeobotanical record recovered since the 1990s suggests
that plant resources played a key role in the territorial expansion
and settling of human populations through Colombia. The data
have enhanced our ability to understand the adaptive strategies of
human groups from ca. 10,500 BP, the Pleistocene/Holocene transition, when archaeological sites start to become more frequent.
Prior to this date archaeological data is scarce, and does not allow
understanding of the role of plants in the ecological strategies of
human groups.
For the sites predating ca. 10,500 BP associated with the Paleoindian stage for Colombian prehistory (Reichel-Dolmatoff, 1965),
human groups were portrayed as specialized megafauna hunters
based on lithic and limited zooarchaeological records (Correal,
 pez, 1999). Traceological analysis done by Nieuwenhuis
1986; Lo
(2002, p. 66) on 56 stone artifacts from zone I of Tequendama
(Sabana de Bogota) dated between ca. 11,000 and 10,000 BP
concluded that 5 (9%) used wood, bone, and dry skin. The analysis
did not suggest any further plant usage for this timeframe.
For the case of Middle Magdalena, the continuity in lithic
technology and the absence of megafauna in the archaeological
record have been used as arguments to doubt the existence of
specialized megafauna hunters in this region of Colombia (Otero
and Santos, 2002). Several lines of evidence suggest that the human groups settling the Middle Magdalena had broader economic
strategies that included shing as well as hunting and gathering
(Otero and Santos, 2002). Traceological analysis on Middle

Magdalena stone tools has not been clear, mainly due to chronological issues (most come from surface collections). Nieuwenhuis
(2002) extracted starch grains from two tools without further
identication. The other tools have evidence of skin and wood work
associated with game butchering and shing, supporting Otero and
Santos (2002) in their rejection of the specialized megafauna
hunters hypothesis and their support for a broad spectrum economy that used the river ecosystems of the Middle Magdalena.
Regardless of the fact that more specialized analysis are required
to be able to draw more precise conclusions, the existing results
support Reichel-Domatoff's (1997 [1965] p. 40e41) idea that, due to
the environmental conditions of Northwest South America, early
inhabitants were likely to practice broad spectrum economies, that
differed from the classic Paleoindian idea. Nonetheless, we are far
from understanding in detail human economic strategies before
ca.10,500 BP due to the scarce amount of data as well as the small
amount of specialized analyses that have been applied on the
existing evidence. The evidence suggest that, even though hunting
was an important strategy for the earliest inhabitants on Northwest
South America, it was not aimed only at big game and it probably
included a wide range of medium and small species, as well as some
shing in the case of the Middle Magdalena.
Starting at about ca.10,500 BP, the general picture of Colombian
prehistory becomes clearer, as the amount of sites and evidence
increase. Between ca. 10,500 and 8000 BP, lithic technology and
archaeobotanical evidence strongly suggest that plant resources
played a very important role in the economic strategies of human
groups settling in the forests (mainly on the Andes) of Northwest
South America.
Aceituno et al. (2013) argue that the increase in site frequency
after the Pleistocene/Holocene transition (ca. 10,500 BP) is associated with the expansion of human populations to the inter-Andean
valleys. This expansion happened at a time of environmental
changes that in broad terms entailed the expansion of tropical
forests (Piperno and Pearsall, 1998). The lithic technology, as well as
the archaeobotanical evidence recovered from several Andean regions (Popayan, Calima, Middle Cauca, Porce River Basin, and
~ a Roja (Middle Caqueta River Basin),
Medellin River Basin) and Pen
suggest that this expansion's success was in part due to the
increasing reliance on tropical plant resources. All these data suggest some form of plant management that likely included selection
and protection.
The evidence of plant usage has led to discussions about cultivation and early domestication. There is some evidence that suggests the idea of some form of cultivation in areas near Calima,
~ a Roja. For Calima, the authors have
Popayan, Middle Cauca, and Pen
interpreted the hoes as indicators of plot preparation for cultivation
(Cardale et al., 1989; Gnecco and Salgado, 1989). For Popayan,
several lines of argument have been used including the pollen of
colonizer plant species along with edible foreign plants that
required human intervention for their geographic dispersion such
as Lagenaria sp., cf. Ipomoea/Manihot (Gnecco and Mora, 1997;
Gnecco, 2003). Gnecco (2003) suggests that Persea americana
may have been domesticated in Popayan based on the size of the
seeds recovered. Despite the fact that the logic of the arguments is
very structured, stronger data are needed. The palynological
interpretation is based on very few plant species, when normally
pollen-based reconstructions take into account many different
kinds of plants to establish ecological relations that can explain
changes or stasis that may or may not be due to anthropic issues.
When pollen is extracted from archaeological sites, it is harder to
determine if percentage changes are due to activities related to
plant cultivation or merely to open spaces for camp sites. It would
be desirable to reexamine some of the preliminary starch grain
identications and compare them with the more robust databases

Please cite this article in press as: Aceituno, F.J., Loaiza, N., The role of plants in the early human settlement of Northwest South America,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.06.027

F.J. Aceituno, N. Loaiza / Quaternary International xxx (2014) 1e8

existing today to be able to conrm genera such as Manihot and


Ipomoea.
For the Middle Cauca, the starch grains recovered from stone
tools supports the idea that access to carbohydrate rich resources
was a central issue in the human adaptation to tropical environments (Piperno and Pearsall, 1998, p. 53). However, unlike Popayan,
for the Middle Cauca more caution is required concerning the
assumption that the archaeobotanical record can be directly
attributed to cultivated plants. For the time frame that concerns us,
the pollen record displays evidence of forest disturbances, but they
cannot be directly assumed as evidence for plot preparation for
plant cultivation. Furthermore, there are no edaphological studies
that may suggest soil preparation for cultivation. Nonetheless, we
consider as a likely the hypothesis suggesting that since the Pleistocene/Holocene transition human settlers were planting wild
plants in areas near their camps, aiming to optimize collection
times, as a strategy to counteract the fact that in tropical environments there is a dispersed pattern of wild food plants, meaning low
return rates (Piperno and Pearsall, 1998).
~ a Roja is very interesting because it has solid
The case of Pen
evidence of plant cultivation reaching to ca. 9300 BP. On the one
hand there are thousands of macrobotanical remains of palm trees
that clearly suggest the intentional use and perhaps the creation of
patches of this resource (Gnecco and Mora, 1997; Morcote et al.,
1998; Gnecco, 2003). On the other hand, the identication of
phytoliths of foreign plants (Cucurbita, Lagenaria siceraria, and
Calathea sp.) introduced to the region from drier environments is
clear evidence for cultivation (Piperno and Pearsall, 1998, p.
204e205).
~ a Roja, archaeologists have argued
For both Popayan and Pen
that plant cultivation was being done in anthropic patches created
for that purpose in the local environments (Piperno and Pearsall,
1998; Gnecco, 2003. This case is supported by the nding of
foreign plants that depended on human transportation and introduction to the environments where the archaeological sites are
found, along with indicators of anthropic alteration of the nearby
campsites. Although the data from Porce River, Middle Cauca, and
Calima require more care with inferences, the evidence presented
suggests some form of local plant usage that strengthens the idea
that people were increasing the carrying capacity of the natural
environments by practicing some form of cultivation since the
Pleistocene/Holocene transition.
5. Conclusions
The recovery of archaeobotanical remains from northern South
American sites dating back to the Pleistocene/Holocene transition
has strongly suggested that plant resources played an important
role on the human settling on tropical forests and on expansion to
the Andean regions.
The data before ca.10,500 BP is not robust enough as to allow us
to have a detailed picture about the lifestyles of the human groups
that entered South America through Panama. The earliest sites from
the Sabana de Bogota are still the reference to understand the
economic strategies of the rst Colombian inhabitants. Regarding
the traditional idea of specialized megafauna hunters associated to
the earliest human groups, nowadays archaeologists are suggesting
that broad spectrum economies more likely resemble the strategies
of the rst peoples. This idea is supported by the scarcity of
megafauna remains (only Tibito has megafaunal remains associated
with anthropic evidence) compared to middle and small game, the
unifacial lithic technology, as well as recent interpretations of
specialized analyses. At present, with no specialized microbotanical
analysis done on the evidence predating 10,500 BP, we cannot asses
the role of plants in the earliest inhabitants' economic strategies.

Between the Pleistocene/Holocene transition and the early


Holocene (ca. 10,500 and 8000 BP), the increasing number of sites,
the characteristics of lithic technology, and the recovered plant
remains, suggest an important change in the economic strategies of
human groups. The number and location of archaeological sites
suggest that human groups were settling in the forests of the subAndean valleys of Colombia (Aceituno et al., 2013). There is also
evidence of movements into other regions, as well as the transportation and introduction of foreign plants that were then brought
~ a Roja (Piperno and Pearsall, 1998).
under cultivation, as at Pen
All the data presented above strongly suggest that plant resources played a key role in the settling of human groups in the
forest environments in the Neotropics during the Pleistocene/Holocene transition. Several authors argue that some plants are
foreign to some of the regions they are found, as evidence of
transportation and cultivation. Even though this practice follows
the behavioral ecology model of increasing the carrying capacity of
tropical environments suggested elsewhere (Piperno and Pearsall,
1998), we think that for Colombian archaeology we need more
lines of evidence as well as detailed specialized analysis to make a
stronger case for widespread cultivation practices starting at the
Pleistocene/Holocene transition.
Excluding Pubenza, Sabana de Bogota, and Middle Magdalena,
all the archaeological sites can be identied with Ranere's (1980)
Tropical Forest Archaic, consisting of behaviors adapted to tropical forest resources, developed in the Neotropics at the beginning
of the Holocene (and in our case extended to the Pleistocene/Holocene transition), that entailed the increasing reliance on plants
without completely forgetting the faunal resources (although absent from the archaeological record). The selective use of some
plants during the Pleistocene/Holocene transition through the early
Holocene preceded the development of horticulture as the main
economic strategy in Northwest South America.
References
Aceituno, F.J., 2001. Ocupaciones Tempranas del Bosque Tropical Subandino en la
Cordillera Centro-Occidental de Colombia (Unpublished PhD dissertation).
Facultad de Geografa e Historia, Universidad Complutense de Madrid, Madrid.
n del Bosque en el Cauca Medio
Aceituno, F.J., Loaiza, N., 2007. Domesticacio
Colombiano entre el Pleistoceno Final y el Holoceno Medio. In: British
Archaeological Reports, International Series 1654. Archaeopress, Oxford.
Aceituno, F.J., Loaiza, N., Delgado, E., Barrientos, G., 2013. The initial human settlement of Northwest South America during the Pleistocene/Holocene transition: synthesis and perspectives. Quaternary International 301, 23e33.
Bray, W., Herrera, L., Schrimpff, M., 1988. Report on the 1984 eld season. ProCalima. Arch
aologisch-ethnologisches Projekt im Westlichen Kolumbien/Sdamerika 5. Periodische Publikation der VereinigungProCalima, Basel, pp. 2e42.
Bush, M.B., Colinvaux, P.A., 1990. A pollen record of a complete glacial cycle from
lowland Panama. Journal Vegetation Science 1, 105e118.
Cano, M., 2004. Los primeros habitantes de las cuencas medias de los ros Otn y
 pez, C., Cano, M. (Eds.), Cambios Ambientales en Perspectiva
Consota. In: Lo
rica, Ecorregio
n del Eje Cafetero Volumen I. Universidad Tecnolo
gica de
Histo
Pereira, Programa Ambiental GTZ, Pereira, pp. 68e91.
Cano, M., 2008. Evidencias precer
amicas en el municipio de Pereira: efectos del
n temprana de los bosques ecuatoriales en el abanico
vulcanismo y colonizacio
nico Pereira-Armenia. In: Lo
pez, C., Ospina, G. (Eds.), Ecologa
uvio-volca
rica: Interacciones Sociedad Ambiente a Distintas Escalas Sociotemporales.
Histo
gica de Pereria-Universidad del Cauca-Sociedad ColombiUniversidad Tecnolo
ana de Arqueologa, Pereira, pp. 84e89.
Cardale, M., Bray, W., Herrera, L., 1989. Reconstruyendo el pasado en Calima
resultados recientes. Boletn del Museo del Oro 24, 3e33.
Cardona, Luis C., Luis, E., Nieto, Jorge, I. Pino, 2007. Del Arcaico a la Colonia. Conn del paisaje y cambio social en el Porce Medio. Informe nal. Unistruccio
versidad de Antioquia-Empresas de Medelln, Medelln. Unpublished.
ctrico estudios de arqueologa preCardona, L.C., 2012. Porce III proyecto hidroele
n del paisaje y cambio social en el
ventiva. Del arcaico a la colonia. Construccio
Porce Medio. Empresas Pblicas de Medelln, Medelln.
Castillo, N., Aceituno, F.J., 2006. El bosque domesticado, el bosque cultivado: un
proceso milenario en el valle medio del rio Porce en el noroccidente Colombiano. Latin American Antiquity 17, 561e578.
Cavelier, I., Rodrguez, C., Herrera, L., Morcote, G., Mora, S., 1995. No solo de la caza
n del bosque amazo
 nico, Holoceno temprano. In:
vive el hombre: Ocupacio

rica
Cavelier, I., Mora, S. (Eds.), Ambito
y ocupaciones tempranas de la Ame

Please cite this article in press as: Aceituno, F.J., Loaiza, N., The role of plants in the early human settlement of Northwest South America,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.06.027

F.J. Aceituno, N. Loaiza / Quaternary International xxx (2014) 1e8

n Erigaie, Instituto Colombiano de Antropologa, Bogota


,
Tropical. Fundacio
pp. 27e44.
nica en Colombia.
Correal, G., 1981. Evidencias Culturales y Megafauna Pleistoce
n de Investigaciones Arqueolo
gicas Nacionales, Bogota
.
Fundacio
Correal, G., 1982. Restos de megafauna asociados a artefactos en la Sabana de
. Caldasia 13 (64), 488e547.
Bogota
nico y el hombre
Correal, G., 1986. Apuntes sobre el medio ambiente pleistoce
 rico en Colombia. In: Bryan, A.L. (Ed.), New Evidence for the Pleistocene
prehisto
Peopling of the Americas. Center for the Study of the Early Man, University of
Main, Orono, pp. 115e131.
 gicas en los Abrigos
Correal, G., van der Hammen, T., 1977. Investigaciones Arqueolo
.
Rocosos del Tequendama. Biblioteca Banco Popular, Bogota
Correal, G., van der Hammen, T., Lerman, J.C., 1966-1969. Artefactos lticos de
abrigos rocosos en el Abra, Colombia. Revista Colombiana de Antropologa 14,
9e53.
Espinal, L.S., 1990. Zonas de vida de Colombia. Universidad Nacional de Colombia,
Medelln.
n Temprana de Bosques Tropicales de Montan
~ a. Editorial
Gnecco, C., 2000. Ocupacio
Universidad del Cauca.
Gnecco, C., 1994. The Pleistocene/Holocene Boundary in the Northern Andes: an
archaeological Perspective. Washington University, Department of Anthropology, Saint Louis (Unpublished PhD dissertation).
Gnecco, C., 2003. Against ecological reductionism: Late Pleistocene huntergatherers in the tropical forests of northern South America. Quaternary International 109e110, 13e21.
Gnecco, C., Aceituno, F.J., 2006. Early humanized landscapes in northern South
America. In: Morrow, J.E., Gnecco, C. (Eds.), Paleoindian Archaeology: a Hemispheric Perspective. University Press of Florida, Gainesville, pp. 86e104.
Gnecco, C., Mora, S., 1997. Late Pleistocene/Early Holocene tropical forest occupa~ a Roja, Colombia. Antiquity 71, 683e690.
tions at San Isidro and Pen
Gnecco, C., Salgado, H., 1989. Adaptaciones precer
amicas en el suroccidente de
Colombia. Boletn del Museo del Oro 24, 35e55.
~ os Antes
Groot, A.M., 1992. Checua: Una Secuencia Cultural entre 8500 y 3000 An
 n de Investigaciones Arqueolo
 gicas Nacionales, Banco de
del Presente. Fundacio
la Repblica, Bogot
a.
Herrera, L., Bray, W., Cardale, M., Botero, P., 1988. Nuevas fechas de radiocarbono
para el precer
amico en la Cordillera Occidental de Colombia. Paper Presented at
the 46th Internainal Congreso of Americanists, Amsterdam.
Herrera, L., Bray, W., Cardale, M., Botero, P., 1992. Nuevas fechas de radiocarbono
para el precer
amico de la Cordillera Occidental de Colombia. In: OrtizTroncoso, O., van der Hammen, T. (Eds.), Archaeology and Environment in Latin
America, Institutvoor Pre- en ProtohistorischeAcheologie Albert Egges van
Giffen. Universiteit van Amsterdam, Amsterdam, pp. 145e163.
~ a, O., 2011. La historia muy antigua del municipio de
Herrera, L., Moreno, C., Pen
gico del aeropuerto
Palestina (Caldas). Proyecto de rescate y monitoreo arqueolo
. Centro de Museos-Universidad de Caldas, Asociacio
 n Aeropuerto del
del cafe
 (2005-2011), Manizales.
Cafe
Hurt, W., van der Hammen, T., Correal, C., 1977. The El Abra Rockshelters, Sabana de
, Colombia, South America. Indiana Universitycc Museum, Bloomington.
Bogota
Occasional Papers and Monographs No. 2.
INTEGRAL, 1997. Arqueologa de Rescate: Va Alterna de la Troncal de Occidente Ro
Campoalegre-Estadio Santa Rosa de Cabal. Informe Final. INTEGRAL S.A. Ministerio de Transporte, Instituto Nacional de Vas, Medelln (Unpublished).
pez, C., 1989. Evidencias paleoindias en el valle medio del ro Magdalena
Lo
 y Remedios, Antioquia). Boletn de
(Municipios de Puerto Berro, Yondo
Arqueologa 4 (2), 3e23.
pez, C., 1995. Dispersio
n de puntas de proyectil bifaciales en la cuenca media del
Lo
ro Magdalena. In: Cavelier, I., Mora, S. (Eds.), Ambito y ocupaciones tempranas
rica Tropical. Fundacio
n Erigaie, Instituto Colombiano de Antropode la Ame
loga, Bogot
a, pp. 73e82.
pez, C., 2008. Landscape Development and the Evidence for Early Human
Lo
Occupation in the Inter-Andean Tropical Lowlands of the Magdalena River.
Colombia.SyllabaPress, Miami.
pez, C., 1999. Ocupaciones tempranas en las tierras bajas tropicales del valle
Lo
-Antioquia. Fundacio
 n de
medio del ro Magdalena: sitio 05-Yon-002 Yondo
gicas Nacionales, Banco de la Repblica. Bogota
.
Investigaciones Arqueolo
n de Medio Ro Caqueta
Llanos, J.M., 1997. Artefactos de molienda en la regio
Amazona Colombiana. Boletn de Arqueologa 12 (2), 3e95.
Marchant, R., Behling, H., Berro, J.C., Cleef, A., Duivenvoorden, J., Hooghiemstra, H.,
Kuhry, P., Melief, R., Schreve-Brinkman, E., van Geel, B., van der Hammen, T., van
Reenen, G., Wille, M., 2002. Pollen-based biome reconstructions for Colombia at
3000, 6000, 9000, 12000, 15000 and 18000 14C yr ago: Late Quaternary tropical
vegetation dynamics. Journal of Quaternary Science 17, 113e129.

lisis polnico en el yacimiento la Pochola: un contexto preMercado, J., 2011. Ana


cer
amico en el ro San Eugenio (Cauca medio, Risaralda eColombia-). Instituto
de Biologa, Universidad de Antioquia.
nica un
Mora, S., 2003. Habitantes tempranos de la selva tropical lluviosa amazo
micas humanas y ambientales. Pittsburgh: Universidad
estudio de las dina
nico de InvestigacionesNacional de Colombia eSede Leticia Instituto Amazo
IMANIUniversity of Pittsburgh, Department of Anthropology, Latin American
Archaeology Reports n. 3.
Mora, S., Gnecco, C., 2003. Archaeological hunter-gatherers in tropical rainforests: a
view from Colombia. In: Mercader, J. (Ed.), Under the Canopy: the Archaeology
of Tropical Rain Forests. Rutgers University Press, New Brunswick, pp. 271e290.
Morcote, R.G., Cabrera, C., Mahecha, D., Franky, C., Cavelier, I., 1998. Las palmas entre
los grupos cazadores recolectores de la Amazonia colombiana. Caldasia 20,
57e74.
Nieuwenhuis, C.J., 2002. Traces on Tropical Tools: a Functional Study of Chert Artifacts from Preceramic Sites in Colombia (PhD thesis from Leiden University).
Archaeological Studies Leiden University No. 9. Faculty of Archaeology, University of Leiden, Leiden.
Otero, H., Santos, G., 2002. Aprovechamiento de recursos y estrategias de movilidad
nicos del valle medio del Magdade los grupos cazadores-recolectores holoce
lena, Colombia. Boletn de Antropologa 16 (33), 100e134.
nicas del can
~o
n del
Otero, Helda, y Gustavo, Santos, 2006. Las ocupaciones prehispa
n, rescate y monitoreo arqueolo
 gico. Proyecto hidroele
cro Porce. Prospeccio
trico Porce III. Obras de infraestructura. Informe nal. Universidad de Antioquia-Empresas de Medelln, Medelln (Unpublished).
ctrico estudios de arqueologa
Otero, H., Santos, G., 2012. Porce III proyecto hidroele
mica de cambio en las sociedades prehisp
preventiva. Dina
anicas de la cuenca
baja del Porce. Empresas Pblicas de Medelln, Pereira.
Pinto, M., 2003. Galindo, un Sitio a Cielo Abierto de Cazadores-Recolectores en la
. Fundacio
 n de Investigaciones Arqueolo
gicas Nacionales,
Sabana de Bogota
Banco de la Repblica, Bogot
a.
Piperno, D., 1985. Phytolith records from prehistoric raised elds in the Calima
n, Colombia. Pro-Calima 4, 37e40.
regio
Piperno, D., 1995. Plant microfossils and their application in the New World. In:
Sthal, P.S. (Ed.), Archaeology in the Lowland American Tropics, Current
Analytical Methods and Applications. Cambridge University Press, pp. 130e153.
Piperno, D., Pearsall, D.M., 1998. The Origins of Agriculture in the Lowland Neotropics. Academic Press, San Diego.
Ranere, A., 1980. Adaptive Radiations in Prehistoric Panama. In: Peabody Museum
Monographs No 5. Harvard University Press, Cambridge, pp. 16e43.
Ranere, A., 2006. The Clovis colonization of Central America. In: Morrow, J.E.,
Gnecco, C. (Eds.), Paleoindian Archaeology a Hemispheric Perspective. University Press of Florida, Gainesville, pp. 69e85.
Ranere, A., 2008. Lower central America. In: Pearsall, D. (Ed.), Encyclopedia of
Archaeology. Academic Press, New York, pp. 192e209.
Reichel-Dolmatoff, G., 1965. Colombia Ancient Peoples and Places. Thames and
Hudson, London.
Reichel-Domatoff, G., 1997. Arqueologa de Colombia: un texto Introductorio, sec.
ond ed. Biblioteca familiar Presidencia de la Repblica, Bogota
 gico en el 
Restrepo, C.A., 2013. Informe de monitoreo arqueolo
area de inuencia del
.
Proyecto de Desarrollo Vial Armenia Pereira Manizales, Autopista de Cafe
 S. A., Pereira (Unpublished).
Instituto Nacional de Vas, Autopistas de Cafe
~ os de Presencia Humana en el P
 n de
Rivera, S., 1991. Neusa 9000 An
aramo. Fundacio
gicas Nacionales, Banco de la Repblica, Bogot
Investigaciones Arqueolo
a.
Rodrguez, C.A., 2002. El valle del Cauca prehisp
anico: procesos socioculturales
ricas del alto y medio Cauca y la costa Pacca
antiguos en las Regiones Geohisto
colombo-ecuatoriana. Departamento de Historia, Facultad de Humanidades
Universidad del Valle, Santiago de Cali.
micos en el alto y medio ro Calima,
Salgado, H., 1988-1990. Asentamientos precera
Cordillera Occidental de Colombia. Cespedesia 16e17 (57e58), 139e162.
~ os de Ocupaciones Humanas en Envigado Antioquia. El
Santos, G., 2010. Diez Mil An
Sitio La Morena. Alcalda de Envigado, Envigado.
Snarkis, M.J., 1979. Turrialba: a Paleo-indian quarry and workshop site in eastern
Costa Rica. American Antiquity 44, 125e138.
n ro Campoalegre, mundo arcaico en la regio
n
Tabares, D., 2004. Fase I: Prospeccio
n de Investigaciones Arqueolo
gicas
del Cauca medio, Colombia. Fundacio
 (unpublished).
Nacionales, Bogota
n: arqueologa
Tabares, D., Rojas, E., 2000. Aportes para una historia en construccio
de rescate en la doble calzada Manizales-Pereira-Armenia. INVIAS-CISAN,
Bogot
a (unpublished).
nico en
Van der Hammen, T., Correal, G., 2001. Mastodontes en un humedal Pleistoce
el valle del Magdalena (Colombia) con evidencias de la presencia del hombre en
el pleniglacial. Boletn de Arqueologa 16, 4e36.

Please cite this article in press as: Aceituno, F.J., Loaiza, N., The role of plants in the early human settlement of Northwest South America,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.06.027

Quaternary International xxx (2014) 1e15

Contents lists available at ScienceDirect

Quaternary International
journal homepage: www.elsevier.com/locate/quaint

Alteration of tropical forest vegetation from the PleistoceneeHolocene


transition and plant cultivation from the end of early Holocene
through middle Holocene in Northwest Colombia
Gustavo Santos Vecino a, *, Carlos Albeiro Monsalve Marn b, Luz Victoria Correa Salas b
a
b

n, Municipio de Envigado, Carrera 45 No. 34A Sur-65, Envigado, Colombia


Secretara de Educacio
n en Palinologa y Paleoecologa de la Universidad Nacional de Colombia, Sede Medelln, Calle 59A No. 63-20, Medelln, Colombia
Grupo de Investigacio

a r t i c l e i n f o

a b s t r a c t

Article history:
Available online xxx

In northwest Colombia, recent studies in the Medelln-Porce River Basin have revealed occupations of
human groups that altered tropical forests from the PleistoceneeHolocene transition (10,000 BP) onward, and that likely implemented small-scale cultivation of wild and domesticated plants from the end
of the early Holocene (8000 BP) until the end of the middle Holocene (3000 BP). These interpretations
are based on archaeological evidence of stone tools for processing plants as well as botanical microfossil
indicators of forest disturbance and plants exploitation and cultivation. In a continental context, the
evidence from these studies are consistent with the idea that in South America, early settlers modied
ecosystems through the selection and exploitation of useful species, thus generating anthropogenic
landscapes beginning as early as the PleistoceneeHolocene transition.
2014 Elsevier Ltd and INQUA. All rights reserved.

Keywords:
Early Holocene
Northwestern Colombia
Medelln-Porce River
Humaneenvironment interaction
Horticulture

1. Introduction
Several authors have suggested that agricultural emerged in South
America as a result of continuous interaction between human groups
and environment that began before the end of the Pleistocene
(Gnecco, 2000; Gnecco and Aceituno, 2004; Aceituno and Loaiza,
2007). Recent studies in sites of the Medelln-Porce River Basin,
located in Cordillera Central of Colombia (Department of Antioquia,
Northwestern Colombia), have revealed new and important information, suggesting human groups began altering the montane humid
forests (between 920 and 2170 m asl) as early as the PleistoceneeHolocene transition (10,000 years ago). These sites are La Morena,
 Valley (upper basin) (Santos, 2010, 2011), and Prilocated in Aburra
mavera I and II, located in the impact area of the Porce III Hydroelectric
Project (lower basin) (Otero de Santos and Santos, 2012).
In these sites, stone toolkits for processing vegetables, such as
chipped stone axes, edge ground cobbles (some of them used as
anvils), handstones and milling stone bases, indicate exploitation
and manipulation of plants, and botanical microfossils indicate
forest disturbance and plant cultivation. This evidence suggests
that from the beginning of the early Holocene, the rst settlers
contributed to favorable conditions in the forests, such as the selection and promotion of edible species, which in turn encouraged
* Corresponding author.
E-mail address: gsantosvecino@yahoo.es (G. Santos Vecino).

the development of horticulture, understood as small-scale plantings containing a range of wild and domesticated plants (Piperno
and Pearsall, 1998, pp. 6e7), from the end of early Holocene
(8000 BP) until the end of middle Holocene (3300 BP).
This paper describes the characteristics of the archaeological
sites mentioned above, and the archaeological and the palaeobotanical evidence obtained. It discusses the hypothesis articulated
above
regarding
forest
alteration
from
the
PleistoceneeHolocene transition that favored the emergence of
horticulture during the eighth millennium BP in the basin of the
Medelln-Porce River, taking into account the discussion of the
emergence of agriculture in South America.
2. Archaeological sites and their features
2.1. Site La Morena
The site La Morena (6 70 59,97800 N; 78 33018,62800 W) is located
 Valley in the upper basin
in the slopes that descend to the Aburra
of the Medelln-Porce River, in the Municipality of Envigado,
a
Department of Antioquia (Santos, 2010, 2011) (Fig. 1). The Aburr
Valley (between 1400 and 2000 m asl) is situated between two
branches of the Cordillera Central, which form the Rionegro High
Plain, to the southeast, and Santa Rosa de Osos High Plain, to the
northwest, both greater than 2000 m asl. The site (Fig. 2) is located
in a colluvial deposit at the foot of the escarpment of the Rionegro

http://dx.doi.org/10.1016/j.quaint.2014.09.018
1040-6182/ 2014 Elsevier Ltd and INQUA. All rights reserved.

Please cite this article in press as: Santos Vecino, G., et al., Alteration of tropical forest vegetation from the PleistoceneeHolocene transition and
plant cultivation from the end of early Holocene through middle Holocene in Northwest Colombia, Quaternary International (2014), http://
dx.doi.org/10.1016/j.quaint.2014.09.018

G. Santos Vecino et al. / Quaternary International xxx (2014) 1e15

Fig. 1. Sites with early occupations in Medelln-Porce River.

High Plain, at 2170 m asl, approximately 2.75 km from the


Medelln River (Fig. 1). The climate of the Aburr
a Valley is generally temperate and humid, with temperatures ranging from 16  C
to 24  C, and average annual rainfall of 1593 mm. However,
ascending the slopes the weather turns cooler and wetter; temperatures on Rionegro High Plain range from 13  C to 18  C and
annual rainfall is 2000e2200 mm. Because of its altitude and
geomorphological position, La Morena is located in a transition

 Valley, characzone between the temperate zone of the Aburra


terized by a vegetation of Premontane Wet forest (Holdridge,
1967; Espinal, 1990), and the cool zone of Rionegro High Plain
characterized by a vegetation of Montane Wet Forest (Holdridge,
1967; Espinal, 1990). In this archaeological site, with an approximate area of 500 m2, 213 m2 were excavated between 2007 and
2008, within an archaeological research project that developed
between 2007 and 2011.

Please cite this article in press as: Santos Vecino, G., et al., Alteration of tropical forest vegetation from the PleistoceneeHolocene transition and
plant cultivation from the end of early Holocene through middle Holocene in Northwest Colombia, Quaternary International (2014), http://
dx.doi.org/10.1016/j.quaint.2014.09.018

G. Santos Vecino et al. / Quaternary International xxx (2014) 1e15

2.2. Sites Primavera I and II

 Valley.
Fig. 2. Site La Morena in Aburra

The archaeological deposit has several components, including


preceramic, early ceramic and late ceramic. The preceramic
component is located in the deepest stratum of the deposit, corresponding to the AB horizon (Fig. 3). In the overlying stratum, A2
horizon, the last ceramic occupation, dating to 3180 40 BP (based
on an associated human burial), disturbed the upper part of the site
deposits mixing late ceramic, early ceramic and the upper part of
the preceramic component. Therefore, this paper refers only to the
unaltered preceramic component, the AB horizon, dating in its
lower part to 10,060 60 and 9680 60 BP and in its upper part in
7080 40 and 4170 50 BP (dates give in the text are uncalibrated,
with some explicit exceptions; for calibrated determinations see
Table 1). In this preceramic component were found stone tools for
plant processing, expedient chipped stone artifacts, and debitage or
waste material, as well as small charcoal fragments.

The sites Primavera I (6 460 894400 N; 78 70 14,79500 W;


1180 m asl) and Primavera II (6 580 35,43300 N; 78 50 25,50300 W;
920 m asl), one kilometer apart, are located in the lower MedellnPorce River Basin, in the Porce III Hydroelectric Project impact area,
Municipality of Anor, Department of Antioquia (Otero de Santos
and Santos, 2012) (Fig. 1). The geomorphology of the lower
Medelln-Porce River Basin (350e1800 m asl) is characterized by a
narrow, deep, V-shaped canyon. The sites are located in colluvial
deposits in the west bank of the Porce River (Fig. 4). Primavera I is
located approximately two km from the river, and Primavera II is
located approximately one km from the river (Fig. 1). In this part of
the basin the temperature varies between 19  C and 24  C, and
rainfall is 1900 to 2500 mm annually, with high humidity (between
80% and 90%). Both sites are in a transition zone between the
Tropical Wet Forest (Holdridge, 1967; Espinal, 1990) of the warm
valley oor, and the Premontane Wet Forest (Holdridge, 1967;
Espinal, 1990) of the temperate valley slopes.
These sites were excavated in 2004 within an archaeological
rescue and monitoring project that developed between 2003 and
2007. Primavera I covers approximately 276 m2, of which 10 m2
were excavated, and Primavera II covers an area of 945 m2, of which
44.5 m2 were excavated.
Both archaeological deposits have preceramic and early ceramic
components. These consist of several anthropogenic strata, formed
by successive layers of rock brought to the site from rocky outcrops
with dark soils rich in organic materials and phosphorus, containing small charcoal fragments, many plant processing tools,
expedient chipped stone artifacts, some formal chipped artifacts
and debitage (Figs. 5 and 6). The preceramic strata date to
7190 40, 7110 40 and 6890 40 BP in Primavera I (horizons
lower A3, A4 and A5), and 7730 170 BP in Primavera II (horizon
2A3) (Table 1). The early ceramic components date to 3300 70 BP

Table 1
Chronology of the earliest sites La Morena and Primavera I and II.
Site

Estrata (horizon)

Laboratory number

Material

Conventional age

2 Sigma Calibration Beta

Cultural association

La Morena
La Morena

Lower AB
Lower AB

Beta-245566
Beta-245564

charcoal
charcoal

10,060 60 BP
9680 60 BP

Preceramic
Preceramic

La Morena
La Morena
Primavera I

Upper AB
Upper AB
A5

Beta-260242
Beta-245565
Beta-105282

charcoal
charcoal
charcoal

7080 40 BP
4170 50 BP
7190 40 BP

Primavera I

A4

Beta-205283

charcoal

7110 40 BP

Primavera I

Lower A3

Beta-205284

charcoal

6890 40 BP

Primavera I

Upper A3

Beta-205285

soil

3300 70 BP

Primavera II
Primavera II
Primavera II

2A3
Lower 2A2
2A1

Beta-205296
Beta-205294
Beta-208247

charcoal
charcoal
charcoal

7730 170 BP
4170 40 BP
3680 40 BP

Primavera II
59

2A1
AB

Beta-205297
Beta-231479

charcoal
charcoal

3650 40 BP
8340 40 BP

61

AB

Beta-231482

charcoal

4650 70 BP

10,030e9360 BC (11,980e11,320 BP)


9270e9110 BC
(11,220e11,060 BP)
9080e9050 BC
(11,030e11,000 BP)
9020e8840 BC
(10,970e10,790 BP)
6020e5890 BC (7970e7840 BP)
2890e2580 BC (4840e4530 BP)
6090e5990 BC
(8040e7940 BP)
6030e5890 BC
(7980e7840 BP)
5840e5710 BC
(7790e7660 BP)
1740e1420 BC
(3690e3380 BP)
7060e6230 BC (9010e8180 BP)
2890e2600 BC (4840e4560 BP)
2190e2170 BC
(4140e4120 BP)
2150e1940 BC (4100e3900 BP)
2140e1910 BC (4090e3860 BP)
7520e7320 BC
(9470e9270 BP)
2630e3300 BC
(5580e5280 BP)
3150e130 BC
(5100e5080 BP)
3210e3180 BC
(5160e5130 BP)

Preceramic
Preceramic
Preceramic
Preceramic
Preceramic
Early ceramic
Preceramic
Early ceramic
Early ceramic

Early ceramic
Preceranic
Early ceramic

Please cite this article in press as: Santos Vecino, G., et al., Alteration of tropical forest vegetation from the PleistoceneeHolocene transition and
plant cultivation from the end of early Holocene through middle Holocene in Northwest Colombia, Quaternary International (2014), http://
dx.doi.org/10.1016/j.quaint.2014.09.018

G. Santos Vecino et al. / Quaternary International xxx (2014) 1e15

Fig. 3. Stratigraphy of La Morena site.

Fig. 4. Site Primavera II in Porce River Lower Basin.

in Primavera I (horizon upper A3), and 4170 40, 3680 40 and


3650 40 BP in Primavera II (horizons 2A1 and 2A2).
3. Archaeological evidence
The preceramic of La Morena and Primavera I and II, located
chronologically in the early and middle Holocene (between 10,000
and 3300 BP) according to the dates indicated, have yielded
numerous examples of a particular type of stone tool, dened as
chipped stone axes (Figs. 7 and 8) (Otero de Santos and Santos,
2012; Santos, 2008, 2010, 2011). These tools, usually with their
ends and sides sharp, were formed from large akes or spalls from
boulders of lapilli-tuff, gabbro and diorites, and are relatively light

(in La Morena the average of 15 complete axes is 412.53 284.89 g,


and in Primavera I and II the average of 187 complete axes is
251.27 145.56 g). Some have polished edges on one end, possibly
as a strategy to facilitate the resharpening and to prolong the useful
life of the tool.
 Valley, chipped stone axes have been found in
In the Aburra
excavations at the preceramic site La Blanquita (upper basin of
Quebrada La Guayabala, Municipality of La Estrella), dating to
7720 50 BP (Botero and Martnez, 2002; Botero, 2008); at the
Casablanca site (Pajarito locality of Medelln), associated with an
early ceramic occupation with rock layers (Nieto et al., 2003); and
at El Ranchito site (western bank of the Medelln River, between the
Municipalities of Itag and La Estrella), also associated with an
early ceramic occupation (Acevedo, 2003). In addition, chipped
stone axes have been found supercially in several sites or places,
, in Medelln
such as the upper basin of the Quebrada La Iguana
Valley (Castillo, 1995); Loma de los Ochoa, municipality of Girardota
(GAIA, 1999); the Pedregal site, upper basin of the Quebrada La
Tablaza, between the Municipalities of Medelln and Itag (G.

Santos, personal observation); and the residential areas of Alamos
del Escobero and La Novena, Municipality of Envigado (Santos,
2006).
Chipped stone axes were also found in two other sites excavated
in the Porce III Area, initially recorded as sites 12 and 100 (Otero de
Santos and Santos, 2012) (Fig. 1), and subsequently excavated and
renumbered as 61 and 59, the rst with a preceramic component
dating to in 8340 40 BP, and the second with an early ceramic
component dating between 4650 70 and 3750 50 BP (Cardona,
2012). Chipped stone axes have also been found in three early sites,
identied as 021, 045 and 107, excavated in the area of inuence of
the Porce II Hydroelectric Project (Fig. 1), in the middle basin of the

Fig. 5. Stratigraphy of Primavera I site.

Please cite this article in press as: Santos Vecino, G., et al., Alteration of tropical forest vegetation from the PleistoceneeHolocene transition and
plant cultivation from the end of early Holocene through middle Holocene in Northwest Colombia, Quaternary International (2014), http://
dx.doi.org/10.1016/j.quaint.2014.09.018

G. Santos Vecino et al. / Quaternary International xxx (2014) 1e15

Fig. 6. Stratigraphy of Primavera II site.

Medelln-Porce River (Castillo et al., 2000), which are dated between 7780 90 and 3650 40 BP, according to the radiocarbon
ages reported and revised by Otero de Santos and Santos (2012). In
all of these sites are archaeological deposits formed basically on
rock layers. In addition, chipped stone axes have been found supercially or isolated from early archaeological deposits in several
sites in Porce III Area, in sites 27, 28, 35, 57 and 89.
Some of the chipped stone axes of Porce II Area, Porce III Area
(Primavera I and II sites) and Aburr
a Valley (La Morena and La
Blanquita sites), have side notches (Fig. 9), which makes them
similar to the so called hoes, that have been found in early
Holocene archaeological deposits in the southwest of the country, Cordillera Occidental, in Sauzalito and El Recreo sites in the
Upper Calima region (Cardale et al. (1989, 1992)), site El Pital in


the Middle Calima (Salgado, 1989, 1995), and in Los Arboles
site
n High Plain (Gnecco and Salgado, 1989). These
in the Popaya
hoes are also found for the same time period in Middle Cauca
region, Cordillera Occidental, in sites El Jazmn, El Antojo,
Guayabito, Campoalegre, where also are commonly found chipped stone axes without notches, similar to those found in Antioquia (Tabares, 2003, 2004; 2012; Tabares and Vergara, 1996;
Tabares and Restrepo, 2003) (the chronology of the dated sites
can be seen in Table 2). These morphological similarities suggest
cultural afnities and interactions between groups living in the
interAndean valleys of the Cordilleras Central and Occidental
during the early Holocene, despite the great cultural diversity
that must have existed due to different local and regional
developments.

Table 2
Chronology of sites of early Holocene in Colombia.
Site

Region

Conventional age BP

Cal BC (OxCal 4,1 95,4%)

Reference

El Jazmn
La Morena

Middle Cauca
Medelln-Porce River High Basin

10,120 70
10,060 60

Aceituno and Loaiza (2007)


Santos (2010, 2011)

Sueva
San Isidro
San Isidro
Tequendama I
Tequendama I
Tequendama I
La Morena

Bogot
a High Basin
Popay
an High Plain
Popay
an High Plain
Bogot
a High Plain
Bogot
a High Plain
Bogot
a High Plain
Medelln-Porce River High Basin

10,090
10,050
10,030
10,025
9990
9740
9680

90
100
60
95
100
1135
60

Sauzalito
Sauzalito
San isidro
Sauzalito
~ a Roja
Pen
~ a Roja
Pen
~ a Roja
Pen
El Jazmn
El Recreo
El Recreo Cancha
~ a Roja
Pen
El Antojo
61
Checua
El Recreo Cancha
Guayabito
El Recreo
El recreo
Checua
021

Upper Calima
Upper Calima
Popay
an High Plain
Upper Calima
Amazon Basin
Amazon Basin
Amazon Basin
Middle Cauca
Upper Calima
Upper Calima
Amazon Basin
Middle Cauca
Medelln-Porce River Low Basin
Bogot
a High Plain
Upper Calima
Middle Cauca
Upper Calima
Upper Calima
Bogot
a High Plain
Medelln-Porce River Middle Basin

9670
9600
9530
9300
9250
9160
9125
9020
8750
8550
8510
8360
8340
8200
8030
7990
7980
7830
7800
7780

100
100
100
100
140
900
250
60
160
60
110
60
40
40
7303
100
120
140
160
80

10,075e9449
10,074e9364
10,030e9360a
10,009e9374
10,027e9312
9861e9322
10,015e9299
10,007e9279
9657e8750
9276e8837
9270e9110a
9299e8782
9260e8722
9221e8634
8786e8294
9119e8221
8614e8244
9127e7211
8338e7836
8256e7549
7711e7498
7935e7197
7539e7344b
7520e7320a
7329e7078c
8256e7549
7056e6768
7191e6592c
7060e6440c
7082e6366c
6270e6425a

Correal (1979)
Gnecco (2000)
Gnecco (2000)
Correal and van der Hammen (1977)
Correal and van der Hammen (1977)
Correal and van der Hammen (1977)
Santos (2010, 2011)
Cardale et al. (1989, 1992)
Cardale et al. (1989, 1992)
Gnecco (2000)
Cardale et al. (1989, 1992)
Cavalier et al., 1995
Cavalier et al., 1995
Cavalier et al., 1995
Tabares and Vergara, 1996
Herrera et al. (1992)
Herrera et al. (2011)
Cavalier et al., 1995
Tabares (2004)
Cardona (2012)
Groot de Mahecha (1992)
Herrera et al. (2011)
Mnera and Monsalve (1996)
Herrera et al. (1992)
Herrera et al. (1992)
Groot de Mahecha (1992)
Castillo et al. (2000)
(continued on next page)

Please cite this article in press as: Santos Vecino, G., et al., Alteration of tropical forest vegetation from the PleistoceneeHolocene transition and
plant cultivation from the end of early Holocene through middle Holocene in Northwest Colombia, Quaternary International (2014), http://
dx.doi.org/10.1016/j.quaint.2014.09.018

G. Santos Vecino et al. / Quaternary International xxx (2014) 1e15

Table 2 (continued )
Site

Region

Cal BC (OxCal 4,1 95,4%)

Reference

Primavera II
045
Campoalegre
El Jazmn

Medelln-Porce River Low Basin


Medelln-Porce River Middle Basin
Middle Cauca
Middle Cauca

7730
7710
7600
7590

170
70
90
60

7060e6230a
6500e6440a
6587e6584b
6492e6490b

n
Nemoco
EL Prodigio
El Pital
021
Primavera I
Primavera I
Tequendama I
El Jazmn
La Morena
021
045
021
021
Primavera I

Bogot
a High Plain
Tolima
Middle Calima
Medelln-Porce River
Medelln-Porce River
Medelln-Porce River
Bogot
a High Plain
Middle Cauca
Medelln-Porce River
Medelln-Porce River
Medelln-Porce River
Medelln-Porce River
Medelln-Porce River
Medelln-Porce River

7530
7370
7310
7240
7190
7110
7090
7080
7080
7080
7080
7040
6940
6890

40
130
140
80
40
40
75
60
40
80
130
80
70
40

6462e6350c
6456e6009c
6444e5973c
6195e5950a
6090e5990a
6030e5890a
6090e5779c
6012e5906b
6020e5890a
6030e5735a
6005e5760a
5980e5735a
5950e5635a
5840e5710a

Otero de Santos and Santos (2012)


Castillo et al. (2000)
Mnera and Monsalve (1996)
Tabares and Vergara, 1996 Mnera
and Monsalve (1996)
Correal (1979)
Rodrguez (1991, 1995)
Salgado (1989, 1995)
Castillo et al. (2000)
Otero de Santos and Santos (2012)
Otero de Santos and Santos (2012)
Correal and van der Hammen (1977)
Aceituno and Loaiza (2007)
Santos (2010, 2011)
Castillo et al. (2000)
Castillo et al. (2000)
Castillo et al. (2000)
Castillo et al. (2000)
Otero de Santos and Santos (2012)

a
b
c

Conventional age BP

Middle Basin
Low Basin
Low Basin

High Basin
Middle Basin
Middle Basin
Middle Basin
Middle Basin
High Basin

Beta 2 Sigma Calibration.


Calib Rev. 5.0.1 (Aceituno and Loaiza (2014)).
Intcal 09.14c.

It has been hypothesized that the hoes, by its form, were used
as digging tools (Cardale et al., 1989, 1992; Salgado, 1989, 1995).
However, the use of replicas in generally rocky mountain soils (such
as the Medelln-Porce River Basin), created a characteristic usewear pattern (fractures and large ake scars at the edges) that
was not observed in archaeological chipped stone axes of Antioquia
(Otero de Santos and Santos, 2012). Based on the experimentation
(replication and use of replicas realized by G. Santos), as well as in
observation of the use-wear on the edges, done during the studies
of La Morena and Primavera I and II, it appears that the chipped
stone axes are not suitable for working hardwoods or digging in
soils, but more suitable for working soft materials, possibly to cut
and peel tubers. However, the small size of some of them and the
presence of blunt edges on several suggests that they could have
served as multifunctional tools to cut, scrape and crush other plants
materials (Otero de Santos and Santos, 2012; Santos, 2008, 2010,
2011). In addition, their considerable concentration in sites of
Porce III Area, 4.3/m2 in Primavera I and 6.3/m2 in Primavera II
(although in deposits of thousands of years), suggests that these
were plant processing tools.

Fig. 7. Chipped stone axes. La Morena (layer AB, 4170e10,060 BP).

Fig. 8. Chipped stone axes from Primavera I (layers A3eA5, 7190e3300 BP) and Primavera II (layers 2A1e2A3, 3650e7730 BP).

Fig. 9. Chipped stone axes with notches from La Morena (layer AB, 4170e10,060 BP).

Please cite this article in press as: Santos Vecino, G., et al., Alteration of tropical forest vegetation from the PleistoceneeHolocene transition and
plant cultivation from the end of early Holocene through middle Holocene in Northwest Colombia, Quaternary International (2014), http://
dx.doi.org/10.1016/j.quaint.2014.09.018

G. Santos Vecino et al. / Quaternary International xxx (2014) 1e15

Fig. 10. Edge ground cobbles from Primavera I (layers A3eA5, 3300e7190 BP) and
Primavera II (layers 2A1e2A3, 3650e7730 BP).

Fig. 11. Hand milling stones from La Morena (layer AB, 4170 e10,060 BP).

Fig. 12. Milling stone bases from Primavera I (layers A3eA5, 3300e7190 BP) and
Primavera II (layers 2A1e2A3, 3650e7730 BP).

 Valley (La Morena,


Moreover, in early sites excavated in Aburra
La Blanquita, and Casablanca), the Porce II Area (021, 045 and 107),
and the Porce III Area (Primavera I and II, 12/61 and 100/59),
chipped stone axes always appear in association with ground stone
tools, such as edge ground cobbles (some of them used as anvils),
handstones, and milling stone bases (Figs. 10e12), in addition to
other quartz specimens, such as cores, akes and irregular fragments with edges suitable for cutting and scraping (and some
plane-convex quartz artifacts in Primavera I and II). For this reason,
the chipped stone axes cannot be regarded as isolated tools, but as
elements of a toolkit for plant processing. To explore this idea,
experimentation with cocoyam (Xanthosoma sagittifolium (L.) Shot)
found that these stone toolkits are suitable for cutting, peeling,
reducing and grinding the tubers to convert them to pulp (Santos,
2011).
Therefore, we propose that chipped stone axes, due to their
frequent association with ground stone tools, are components of a
technological organization, understood as the selection and
integration of strategies for making, using, transporting and discarding tools and the materials needed for their manufacture and
maintenance (Nelson, 1991, p. 57). A technological organization
such as the one found in the Medelln-Porce River Basin, is characteristic of groups with restricted mobility (as indicated by the
weight of toolkits and the rock accumulations) who depended
largely on plant foods for their subsistence, along with hunting and
shing (Santos, 2008, 2010, 2011). The low density of early sites in
the lower Medelln-Porce River Basin (only four early sites were
found in 1754 systematically prospected hectares in the nonurbanized Porce III Area, and only three early sites were found in
1119 systematically prospected hectares in the non-urbanized
Porce II Area) suggests a social organization based on small
groups that from several base-camps exploited the resources of
vast territories through logistical movements, represented by
sites with chipped stone axes isolated from early archaeological
deposits (Santos, 2008, 2010, 2011; Otero de Santos and Santos,
2012).
Unfortunately, due to the acidity of soils in the Medelln-Porce
River Basin, organic materials such as faunal remains were normally not preserved. However, exceptional faunal remains were
recovered at one of the sites in the Porce II Area (site 21), dating
between ca. 7000 and ca. 6000 BP. The majority of the faunal remains belong to armadillo (Dasypus novemcinctus), two toed sloth
species (Choloepus hoffmanni and C. didactylus), lowland paca or
spotted paca (Agouti paca), black agouti (Dasyprocta fuliginosa), and
porcupine (Coendou prehensilis). Also represented were red brocket
deer (Mazama americana) and remains of unidentied primates
(Simiformes), carnivores, birds, sh and reptiles (Castillo et al.,
2000). In addition, two quartz projectile points found in the same
site 021, associated with the preceramic occupation dating between
ca. 6000 BP and ca. 5000 BP (Castillo et al., 2000), could have been
used in hunting activities. This assemblage of faunal materials and
the projectile points implies that early groups (preceramic and
early ceramic) of the Medelln-Porce River Basin combined the
collection and cultivation of calorie-rich plant foods, with hunting
and shing for protein-rich animal food, which indicates a broadspectrum subsistence. Therefore, it is possible, that the expedient
quartz artifacts accompanying stone toolkits for plants processing,
and the plane-convex artifacts of Porce III Area, were employed in
preparation of animals, and the manufacture of wooden objects for
hunting and shing. In any case, the wide dispersion of the chipped
 Valley (Fig. 13) shows that groups with
stone axes in the Aburra
technology for plant processing, characteristic of the early and
middle Holocene (10,000e3300 BP), occupied a wide range of
 Valley. These early groups of the Medelln-Porce
habitats in Aburra
River Basin could also have exploited the resources of the cool ora

Please cite this article in press as: Santos Vecino, G., et al., Alteration of tropical forest vegetation from the PleistoceneeHolocene transition and
plant cultivation from the end of early Holocene through middle Holocene in Northwest Colombia, Quaternary International (2014), http://
dx.doi.org/10.1016/j.quaint.2014.09.018

G. Santos Vecino et al. / Quaternary International xxx (2014) 1e15

Fig. 13. Sites with chipped stone axes in Aburra Valley.

of the Cordillera Central, where there were also base-camps as


indicated by the preceramic site El Pedrero with similar stone
toolkits, located in the Rionegro High Plain (Municipality of El
Carmen de Viboral), approximately 27 km from the Aburr
a Valley
(Fig. 1), and dated between 6660 100 and 4510 80 BP (Botero
and Salazar, 1998).
Despite regional differences in expedient aked stone artifacts
and local raw material types, stone toolkits for plant processing

(chipped stone axes or hoes, edge ground cobbles, handstones,


milling stone bases, and in some cases anvils and pounders) are also
found in most sites of the early Holocene located in montane and
lowland forests that have been identied in Colombia, especially in
the temperate zone of the Cordillera Central and Cordillera Occidental, in the southwest section of the country, and in the Middle
Cauca region. These sites are: San Isidro, in the Popayan High Plain,
~ a Roja, in the middle Caqueta
10,050e9530 BP (Gnecco, 2000); Pen

Please cite this article in press as: Santos Vecino, G., et al., Alteration of tropical forest vegetation from the PleistoceneeHolocene transition and
plant cultivation from the end of early Holocene through middle Holocene in Northwest Colombia, Quaternary International (2014), http://
dx.doi.org/10.1016/j.quaint.2014.09.018

G. Santos Vecino et al. / Quaternary International xxx (2014) 1e15

Fig. 14. Starch type. 40 (~14  11 mm). La Morena (layer AB, 4170e10,060 BP).

Fig. 15. Phytoliths type. 40. a) Braquiolita (~21  32 mm); b) Braquiolita (~19  22 mm); c) Globulita (~6e8 mm). La Morena (layer AB, 4170e10,060 BP).

River, in the Amazon basin, 9250e9125 BP (Cavalier et al., 1995);


Sauzalito and El Recreo in the Upper Calima region, Cordillera
Occidental, 9670e9300 BP, and between 8750e7830 BP respectively (Bray et al., 1988; Herrera et al., 1992); El Pital, in the Middle
Calima region, 7310 BP (Salgado, 1989, 1995); El Prodigio, in the
eastern slope of the Cordillera Central, Municipality of Chaparral
(Tolima Department), 7370 BP (Rodrguez, 1991, 1995); and several
sites of Middle Cauca region, such as El Jazmn, 9020 and 7599 BP,
Guayabito, 7990 BP, and Campoalegre, 7600 BP (Mnera and
Monsalve, 1996; Tabares and Vergara, 1996; Aceituno, 2003;
Tabares, 2004, 2012) (the detailed chronology of these sites can
be seen in Table 2). Aceituno et al. (2013) argued that in Colombia,
the lithic technology in the early Holocene (10,000e8000 BP) in
middle Cauca River Basin, middle Porce, and Calima River Basin,
with a high frequency of grinding stones, axes and adzes, along
with several tool types made on akes, seems to be mainly oriented
to procure plant resources.
, where several sites have been
In the high plain of Bogota
identied with occupations of the early Holocene such as
 n and Checua, that are characterized
Tequendama, Sueva, Nemoco
by expedient chipped stone tools, but not chipped axes or hoes.
However, boulders, pounders, and edge ground cobbles are reported (Correal and van der Hammen, 1977; Correal, 1979; Groot de
Maecha, 1992), generally associated with gathering activities
(Correal and van der Hammen, 1977; Correal, 1979), and more
specically to processing seeds and tubers (Groot de Maecha, 1992).

Fig. 16. Pollen Dioscorea type. 40 (~38e40 mm). La Morena (layer upper AB, 7080
40 BP).

Furthermore, although expedient artifacts have been associated


with the processing of faunal resources, this traditional assumption
is being revised, and it is now suggested that this technology was
oriented toward broad-spectrum subsistence, even with an
emphasis on plant exploitation (Aceituno et al., 2013).
Similar grinding tools (edge ground cobbles and milling stone
bases) have been found in Panama, at rockshelter Carabali (Central
Pacic slopes), in layers dating to 8000e7000 BP and in Ecuador, at
Las Vegas site (on the Pacic Coast), in occupations dating to
9800e6600 BP (Stothert, 1985, 1988; Ranere and Cooke, 1995;
Piperno and Pearsall, 1998). This wide distribution of plant processing tools during the early Holocene suggests the existence of a
technological horizon in Colombia and throughout northwest
South America (including Panama), and that the exploitation of
plants was an important activity since the beginning of the
Holocene.
4. Palaeobotanical evidence
 Valley), the paleobotanical
In La Morena (Envigado, Aburra
analysis (pollen, phytoliths and starches) of the soil column
extracted from a prole of the archaeological deposit, the paleobotanical analysis of sediments recovered from stone tools, as well
as the seed identications were made by biologist C. A. Monsalve
(2008a, 2008b, 2009). In Primavera I and II, pollen analysis of soil
samples from proles of archaeological deposits, and starch and
phytolith analyses of sediments recovered from stone tools were
made by Correa (2005).
The treatment of pollen samples proceeded according to the
standard method proposed by Faegri and Iversen (1975), with a
nal application of acetolysis proposed by Erdtman (1969), and
modications suggested by Fonnegra (1989). Taxonomic identication of pollen was carried out using specialized catalogs of
Roubick and Moreno (1991), Murillo and Bless (1974, 1978), Herrera
and Urrego (1996), Salomons (1989), Punt and Blackmore (1994)
and Erdtman (1969, 1971). Treatment of phytoliths proceeded according to procedures described by Zucol and Osterieth (2002), and
the taxonomic identication of phytoliths was done based on their
morphometric characters according to the Bertoldi de Pomar

Please cite this article in press as: Santos Vecino, G., et al., Alteration of tropical forest vegetation from the PleistoceneeHolocene transition and
plant cultivation from the end of early Holocene through middle Holocene in Northwest Colombia, Quaternary International (2014), http://
dx.doi.org/10.1016/j.quaint.2014.09.018

10

G. Santos Vecino et al. / Quaternary International xxx (2014) 1e15

Fig. 17. Pollen diagram of La Morena site.

morphological classication (1975), in which 13 morphotribes and


70 variants are established, and the botanical origin recorded in
Flores (1999). It also took into account the Monsalve (2000) and
Piperno (1998) catalogs, and the personal collection of permanent plates of pollen and starches of L.V. Correa. Treatment of starch
proceeded according to standard protocols (Piperno et al., 2000,
2009; Pearsall et al., 2004) and the taxonomic identication of
starches was done based on their morphometric characters acnez (2009, 2011).
an-Jime
cording to Pag
In La Morena, in the pollen sample of the lower horizon AB,
closely associated with the date 9680 60 BP (obtained in the
same excavation square of 1 m2 and at the same depth), and
stratigraphically associated with the date 10,060 60 BP, were
identied starch fragments (or amyloplasts that clearly show the
characteristic morphology of starch) (Fig. 14). Also recovered from
the same sample were phytoliths types such as Braquiolita, Flabelolita and Globulita (bulliform, cuneiform bulliform, polyhedral
bulliform and fanshaped) (Fig. 15) typical of woody plants, such as
shrubs and trees of the family Melastomataceae Juss., Ericaceae
Juss. and Myrtaceae Juss., which are considered as natural or
anthropogenic disturbance indicators of altered forest areas from
the time of PleistoceneeHolocene transition. Also notable is the
presence of pollen indicators of open vegetation such as Dioscorea

L. (Fig. 16), Taraxacum F.H. Wigg., Asteraceae Martinov, and Poaceae Barhart, and vegetation related to crops (Amaranthaceae
Juss.) throughout the preceramic horizon or AB horizon (Fig. 17)
suggesting forest disturbance around the site. In addition, the
occurrence of signicant percentages of black and brown amorphous microfragments of charcoal and burned plant tissues, resins
and fungi spores in the pollen samples of the preceramic horizon
(as well as the presence of carbon fragments in this horizon),
would indicate activities related to the use of vegetable materials
for energy purposes.
In the sample pollen of the upper horizon AB, closely associated with a date of 7080 40, were identied pollen of Zea mays L.
(Fig. 18) and Dioscorea L., and also starch grain and phytolith indicators of forest disturbance. Furthermore, pollen of Z. mays L.
retrieved from an axe found in the same square and the same level
as the charcoal sample dated in 7080 40 BP. Also, at the horizon
AB, pollen of Z. mays L. and Phaseolus L. were retrieved from a
milling stone base, and pollen of Persea Mill. Was retrieved from
several axes, and. seeds of Arecaceae Bercht and Presl and Phaseolus L., were identied. However, the chronology could not be
determined because the tools and seeds were not closely associated with dates (the description of each taxon can be seen in
Table 3).

Table 3
Description of taxa identied in La Morena and Primavera I and II.
Site

Taxa

Microfossil

Description

La Morena

Type Zea mays L. (Poaceae)

Pollen

La Morena

Type Dioscorea L (Dioscoreaceae)

Pollen

La Morena

Type Phaseolus L. (Fabaceae)

Pollen

La Morena

Type Aiphanes Willd (Arecaceae)

Pollen

Monad; heteropolar-radiosymmetric; monoporat; exina tectate; 1.0e2.0 m thick; exina psilate/


scabrate; amb circular; grain ~70  75 m.
Monad; isopolar-radiosymmetric; exina tectate to intectate; 1.5e2.0 m thick; sulci as long grain;
grain prolate to subprolate ~41e43 m of long.
Monad; Isopolar-radiosymmetric; triporate; exina semitectate, exina microreticulate ~2.0 m thick;
amb circular; grain oblate-spheroidal; ~52e51  53e55 m.
Monad; isopolar-radiosymmetric; mosulcate, sulci short, inconspicuous; exina tectate <1.0 m thick;
micro-echinate, echini ~<1.0 m long; grain oblate-spheroidal, 28e31 m.

La Morena

Type Persea Mill (Lauraceae)

Pollen

Please cite this article in press as: Santos Vecino, G., et al., Alteration of tropical forest vegetation from the PleistoceneeHolocene transition and
plant cultivation from the end of early Holocene through middle Holocene in Northwest Colombia, Quaternary International (2014), http://
dx.doi.org/10.1016/j.quaint.2014.09.018

G. Santos Vecino et al. / Quaternary International xxx (2014) 1e15

11

Table 3 (continued )
Site

Primavera
Primavera
Primavera
Primavera

Taxa

I-II
I-II
I-II
I-II

Primavera I-II

Type
Type
Type
Type

Microfossil

Zea mays L. (Poaceae)


Phaseolus L. (Fabaceae)
Zea mays L. (Poaceae)
Phaseolus L. (Fabaceae)

Type Zea mays L. (Poaceae)

Pollen
Pollen
Phytoliths
Starch

Starch

Description
Monad; apolar-asymmetric; inaperturate; intectate; sexina echinate; exina <0.5 m thick; grain
spheroidal, ~36e41 m.
Monad; spheroidal/elliptical; monoporat; exina ne escabrate; grain 85 and 98 m.
Monad; spheroidal; aperture 4e6 m; asymmetric pores without ring; enyne reticulate, large lumen.
Wavy-tops and rufe tops rondels.
Grains have a central hilum with concentric lamellae with an average size between 12 and 17
microns, its shape is round to oval. Positive test for iodine (Lugol) and have the maltese cross with
crossed nicols.
Grains have a central hilum to form elongated or groove sometimes leasura central grooves
projecting towards the edges of starch grain form a multi-pointed star, its oval shape. Positive test
for iodine (Lugol) and have the Maltese cross with crossed nicols.

Fig. 18. Pollen Zea mays type. 40 (~76e78 mm). La Morena (layer upper AB, 7080 40
BP).

In Primavera II (Porce River Lower Basin), in the pollen sample


from lower horizon 2A3, dating to 7730 170 BP, pollen of Z. mays
L. (Fig. 19) and Phaseolus L. (Fig. 20) were identied. Associated with
the same horizon and the same date were starch grains and phytoliths of Z. mays L. (Fig. 21) retrieved from a milling stone base and
from an axe respectively. Also, in pollen sample from lower horizon
2A2, associated with the early ceramic occupation dating to
4170 40 BP (in which the toolkits for plant processing of the

preceramic occupation continues), pollen of Phaseolus L., starch and


phytoliths of Z. mays L. were retrieved from stone tools. In Primavera I (Porce River Lower Basin), associated with the pollen sample
of the lower horizon 2A3, dating to 6890 40 BP, were phytoliths of
Z. mays L. on two axes. Furthermore, in Primavera I and II in the
preceramic and early ceramic horizons indicators of forest disturbance occur beginning in the eighth millennium BP in the form of
border elements (Piper L. and Rubiaceae Juss.), pioneers elements
(Asteraceae Martinov and Gramineae Barhart) and open vegetation
(Asteraceae Martinov, Gramineae Barhart and Cyperacea Juss.)
(Fig. 22).
In synthesis, in the Medelln-Porce River Basin microfossils of
edible plants have been identied: 1) Z. mays L. (maize) dating to
7730 170 BP, 7080 40 BP, 6890 40 BP and 4170 40 BP; 2)
Dioscorea L. (yam) dating to 10,060 60, 9680 60 BP and
7080 40 BP; and 3) Phaseolus L. (bean) dating to 7730 170 BP
(Table 4). Also present are microfossils of forest disturbance indicators throughout the preceramic horizon in La Morena, dating
between 10,060e9680 BP and 4170 BP, and throughout the preceramic horizons in Primavera I and II, dating between 7730 BP and
4170 BP.

Fig. 19. Pollen Zea mays type. 40. Primavera II (layer 2A3, 7730 170 BP).

Fig. 20. Pollen Phaseolus L. 40. Primavera II (layer 2A3, 7730 170 BP).

Please cite this article in press as: Santos Vecino, G., et al., Alteration of tropical forest vegetation from the PleistoceneeHolocene transition and
plant cultivation from the end of early Holocene through middle Holocene in Northwest Colombia, Quaternary International (2014), http://
dx.doi.org/10.1016/j.quaint.2014.09.018

12

G. Santos Vecino et al. / Quaternary International xxx (2014) 1e15

Table 4
Crop plant occurrence, chronology and provenance in La Morena and Primavera I
and II
Site

Horizon

Primavera II
Primavera II
Primavera II
La Morena
La Morena
Primavera I

2A3
2A3
2A3
AB
AB
Lower 2A3

Age
7730
7730
7730
7080
7080
6890

Primavera I
Primavera I

Lower 2A2
Lower 2A2

4170 40
4170 40

Primavera I
La Morena

A3
Lower AB

La Morena
Primavera I

Upper AB
2A3

3300
10060
9680
7080
7730

170
170
170
40
40
40

Provenance

Crop

Soil sample
Soil sample
Milling stone base
Soil sample
Chipped stone axe
2 Chipped stone
axes
Milling stone hand
2 Chhipped stone
axes
Milling stone hand
Soil sample

Maize
Maize
Maize
Maize
Maize
Maize

70
60
60
40 Soil sample
170 Soil sample

(pollen)
(phytoliths)
(starch)
(pollen)
(pollen)
(phytoliths)

Maize (starch)
Maize (phytoliths)
Maize (starch)
Dioscorea (pollen)
Dioscorea (pollen)
Phaseolus (pollen

5. Discussion
In La Morena, the presence of pollen of open vegetation and
vegetation related to crops, starch grains and phytoliths typical of
shrubs, and microfragments of charcoal and burned plant tissues
(as well as the presence of carbon fragments in soil), resins and
fungi spores are indicative of forest disturbance from the PleistoceneeHolocene transition (10,060e9680 BP) onward. However,
there is no information to measure the disturbance, or to establish
if deforestation was anthropogenic or natural. Nonetheless,
considering that these indicators are in a cultural context characterized by strata with strong anthropogenic inuence and stone
toolkits for plant processing, which presumes the exploitation and
manipulation of plants (stimulus propagation of plants with specic characteristics), it is reasonable to say that there must have
been some anthropogenic alteration of the forest.
The introduction of maize (Z. mays L.) in Medellin-Porce River
Basin must have occurred by the eighth millennium BP, according
to plant microfossils (phytoliths, pollen, and starch) and the earliest
dates (7730 170 BP, 7080 40 BP and 6890 40 BP). This is
acceptable considering that maize is present in Central Balsas River
Valley of Mexico at least by 7920 40 BP (Piperno et al., 2009;
Piperno, 2011) and possibly around 9000 cal BP based on molecular evidence (Matzuoko et al., 2002); is dispersed in lower Central
America by 7600 BP (Piperno, 2011); and is introduced in the
Middle Cauca region (Colombia, about 200 km south of the
Medellin-Porce River Basin) around 7000 BP (Aceituno and Loaiza,
2014).
At the same time (7730e6890 BP), pollen of Dioscorea L. (yam)
in La Morena, and Phaseolus L. (beans) in Primavera II indicate the
local availability of these plants. However, it is not possible to know
if some of these plants were cultivated or domesticated. A similar

situation occurs in the nearby Middle Cauca region, where at least


one species of Dioscorea was being ground and chopped with stone
tools around 7600 BP, and wild Phaseolus species were being processed with stone tools around 8700 BP (Aceituno and Loaiza,
2014). Additionally, in site 021 of Porce II Area pollen from Manihot spp. dating between ca. 6500 and ca. 6000 BP, falls within the
range of ca. 7000 and ca. 5000 BP set for the presence of Manihot
spp. in the Middle Cauca region (Aceituno and Loaiza, 2014).
In the basin of Medellin-Porce River, stone toolkits for plants
processing, associate with indicators of forest disturbance, suggest
alteration of tropical forests from the PleistoceneeHolocene transition onward. Evidence indicates the presence and consumption of
maize, accompanied by other plants such as beans and yams,
suggest preparation of plots for planting and harvesting domesticated and wild plants, or horticulture, beginning in the eighth
millennium BP. This is consistent with information from Middle
Cauca region, where there are indicators of forest disturbance and
exploitation and manipulation of plants from the PleistoceneeHolocene transition, and where horticulture, based on the
presence of maize and possibly manioc, was developed from the
beginning of Middle Holocene (7000 BP) (Aceituno and Loaiza,
2014).

6. Conclusions
In the Medelln-Porce River Basin (Northwest Colombia), there
are several excavated sites (La Morena, La Blanquita, Casablanca,
Primavera I, Primavera II, 12/61, 100/59, 021, 045 and 107) with
preceramic and early ceramic deposits, dating between
10,060e9860 BP and 3300 BP, characterized by stone toolkits for
plant processing that include chipped stone axes, edge ground
cobbles, handstones and milling stone bases as well as expedient
chipped artifacts. This sites document the exploitation and
manipulation of plants from the PleistoceneeHolocene transition
until the end of the middle Holocene. Supporting this interpretation, in the preceramic component of La Morena site, dating between 10,060 and 4170 BP, there are microfossils recovered from
closely associated sediments indicating the presence of open and
shrub vegetation, that suggest forest alteration from PleistoceneeHolocene transition, although there is no information on
whether there was deforestation or forest clearance. The presence
of microfragments of charcoal and burned plant tissues resins and
fungi spores, as well as charcoal in soil deposits, indicate forest
alteration. Furthermore, in the preceramic components of La Morena, and Primavera I and II, between 7730 BP and 6890 BP, associated with indicators of forest disturbance, microfossils
(phytoliths, pollen, and starch) retrieved from stone tools or closely
associated sediments indicate the presence of maize (Z. mays L.), a
domesticated plant, and other plants such beans (Phaseolus L.) and
yam (Dioscorea L), suggesting horticulture, or small-scale plantings

Fig. 21. a) Starch Zea mays L. 40. Retrieved from milling stone base. Primavera II (layer 2A3, 7730 170 BP); b y c) Phytoliths Zea mays. 40. Primavera II (layer 2A3,
7730 170 BP).

Please cite this article in press as: Santos Vecino, G., et al., Alteration of tropical forest vegetation from the PleistoceneeHolocene transition and
plant cultivation from the end of early Holocene through middle Holocene in Northwest Colombia, Quaternary International (2014), http://
dx.doi.org/10.1016/j.quaint.2014.09.018

G. Santos Vecino et al. / Quaternary International xxx (2014) 1e15

13

Fig. 22. Pollen diagrams of Primavera I and II sites.

containing a range of wild and domesticated plants, from eight


millennium BP until end of middle Holocene in Medelln-Porce
River Basin.
This information is similar to that obtained in the neighboring
region of the Middle Cauca, where archaeobotanical evidence
suggests that human groups exploited and manipulated plant since
the PleistoceneeHolocene transition and cultivated plants since the
middle Holocene. The archaeobotanical information of Central and
South America indicates the onset of food production and the
emergence of plant cultivation. The Medelln-Porce River Basin, as
the Middle Cauca, is an important region to study and understand
the emergence and development of agriculture in northern South
America.
Moreover, in the context of tropical forests of America, there is
evidence that soon after the end of the Pleistocene, about
11,400 cal. BP, people settled in their territories, staying longer or
more frequently returning to specic sites (as suggested by Primavera I and II sites), and frequently altered and manipulated their
environments by creating clearings in forests or burning them. The
toolkits developed at this time are evidence that for the rst time
the exploitation of plants was as important an economic strategy as
hunting had been (Piperno et al., 2009; Piperno, 2011).

Therefore, in terms of adaptive strategies, as has been documented in Colombian regions, including Middle Cauca, Middle
n High Plain, Amazon Basin, and
Porce, Medelln River, Popaya
Calima River Basin, lithic technology and archaeobotanical data
indicate a strong focus on plant resources in a time of climate
change in the Neotropics marked by the expansion of tropical forests (Aceituno et al., 2013). Additional evidence shows forest
alteration, such as forest clearing, burning and cultural selection of
key resources increases the forest carrying capacity (Aceituno et al.,
2013). These data have given rise to the hypothesis that these early
hunter-gatherers altered the local ecosystems as part of their
strategies for expansion and adaptation (Aceituno et al., 2013).
Also, in the context of Central and South America, it has been
argued that the archaeobotanical information from various localities, especially in Central Balsas Valley southwestern Mexico, the
central Pacic and western regions of Panama, the sub-Andean and
Premontane zones of the Cauca and Porce valleys in Colombia, the
~ a Valley of
Colombian Amazon, southwestern Ecuador and Zan
northern Peru, indicates that food production began between
11,000 and 7600 BP (Piperno, 2011). Agricultural intensication,
indicated by the number of sites and signicant increase in density
of artifacts, began before 7000 BP when people became slash and

Please cite this article in press as: Santos Vecino, G., et al., Alteration of tropical forest vegetation from the PleistoceneeHolocene transition and
plant cultivation from the end of early Holocene through middle Holocene in Northwest Colombia, Quaternary International (2014), http://
dx.doi.org/10.1016/j.quaint.2014.09.018

14

G. Santos Vecino et al. / Quaternary International xxx (2014) 1e15

burn cultivators integrating planting with collecting and hunting,


thereby taking an important step towards full-scale agriculture
(Piperno, 2011).
On the other hand, these results are consistent with the approaches of several researchers in the sense that, since the end of
the Pleistocene and during the early Holocene, the Andean forests
of the Neotropics were altered by hunteregatherers to stimulate
the growth of edible plants, through the creating of clearings by
deforestation or burning, producing changes in the distribution of
plant and animal species, and fostering the creation of favorable
environments for the cultivation and domestication of plants.
Agriculture, therefore, should be seen as the result of a long process
of interaction between humans and the environment (Gnecco,
2000; Gnecco and Aceituno, 2004; Aceituno and Loaiza, 2007).
The archaeobotanical data of several regions in Colombia, among
which are Middle Cauca and Medelln-Porce River Basin, document
the importance of plants and the alteration of tropical forests as an
adaptive strategy for the early inhabitants of these regions, questioning the stereotype of highly mobile hunteregatherers with a
strong orientation toward hunting animals, which has dominated
most studies of early colonization of America (Gnecco and
Aceituno, 2004; Aceituno et al., 2013).
References
Aceituno, F.J., 2003. Interacciones toculturales en el Cauca Medio durante el Hol
oceno temprano y medio. Arqueologa del Area
Intermedia 4, 89e113.
n del bosque en el Cauca Medio
Aceituno, F.J., Loaiza, N., 2007. Domesticacio
colombiano entre el Pleistoceno Final y el Holoceno medio. British Archaeological Reports. In: International Series 1654. Archeopress, Oxford.
Aceituno, F.J., Loaiza, N., Delgado-Burbano, M.E., Barrientos, G., 2013. The initial
human settlement of Northwest South America during the Pleistocene/Holocene transition: synthesis and perspectives. Quaternary Science Reviews 86,
49e62.
Aceituno, F.J., Loaiza, N., 2014. Early and middle Holocene evidence for plant use
and cultivation in the Middle Cauca River Basin, Cordillera Central (Colombia).
Quaternary International 301, 23e33.
.
Acevedo, J.L., 2003. Aldea y sistemas de canales del siglo III DC en el Valle de Aburra
gico El Ranchito. Predios del Sur S. A. Medelln (UnPlan de Manejo Arqueolo
published results).
Bertoldi de Pomar, H., 1975. Los silicotolitos: sinopsis de su conocimiento.
Darwinian 19, 173e206.
Botero, S., Salazar, C., 1998. El Pedrero. Evidencias de antiguos especialistas en el
Municipio de El Carmen de Viboral, Antioquia-Colombia. Boletn de Antropologa 12 (33), 168e195.
Botero, S.H., 2008. Ocupaciones tempranas en el Valle de Aburr
a. Sitio La Blanquita.
 rica. Interacciones Sociedad-Ambiente a Distintas Escalas
In: Ecologa Histo
pez y Guillermo Ospina compiladores. Universidad
Socio-Temporales. Carlos Lo
gica de Pereira, Universidad del Cauca y Sociedad Colombiana de
Tecnolo
 n.
Arqueologa, Pereira, pp. 239e248. CD versio
Botero, S.H., Martnez, L.H., 2002. Ocupaciones tempranas en Valle de Aburr
a. Sitio
n Cultura y Artes Antioquia (Unpublished
La Blanquita. Fondo Mixto. Promocio
results).
Bray, W., Herrera, L., Schrimpff, M., 1988. Report on the 1984 eld season. ProCalima. Projeckt im WestlichenKolumbien/Sdamerika 5, 2e47.
Cardale de Scrhimpff, M., Bray, W., Herrera, L., 1989. Reconstruyendo el pasado en
Calima. Resultados recientes. Boletn Museo del Oro 24, 3e34. Banco de la
Repblica.
~ os de historia en el
Cardale, M., Bray, W., Herrera, L., 1992. Calima: diez mil an
n Pro Calima, Santafe
 de Bogota
.
suroccidente de Colombia. Fundacio
 n del paisaje y cambio
Cardona, L.C., 2012. Del Arcaico a la Colonia. Construccio
ctrico. Estudios de
social en Porce Medio. In: Porce III. Proyecto hidroele
arqueologa preventiva. Empresas Pblicas de Medelln E.S.P. Universidad de
Antioquia, Medelln, pp. 230e391.
gico en el Valle de Aburra
. Boletn de
Castillo, N., 1995. Reconocimiento arqueolo
Antropologa 9 (25), 49e90.
rrez, J.,
Castillo, N., Aceituno, F.J., Cardona, L.C., Garca, D.P., Pino, J.I., Forero, J.C., Gutie
~ os de historia en el Valle Medio del ro
2000. Entre el bosque y el ro: 10.000 an
Porce. Universidad de Antioquia. Empresas Pblicas de Medelln, Medelln
(Unpublished results).
lo de caza
Cavalier, I., Rodrguez, C., Herrera, L.F., Morcote, G., Mora, S., 1995. No so
n del bosque amazo
 nico, Holoceno temprano. In:
vive el hombre. Ocupacio

s, Mora, Santiago (Eds.), Ambito
Cavalier, Ine
y ocupaciones tempranas de la
rica tropical. Instituto colombiano de Antropologa, Colcultura, Fundacio
n
Ame
 de Bogot
Erigaie, Santafe
a, pp. 27e44.
lisis en sedimentos y artefactos arqueolo
 gicos hallados en el
Correa, L.V., 2005. Ana

ctrico Porce III-Obras de Infraestructura.
area de inuencia del Proyecto Hidroele

Empresas Pblicas de Medelln (EE PP M). Universidad de Antioquia (Unpublished results).


gicas en los abrigos rocosos de Nemoco
n
Correal, G., 1979. Investigaciones arqueolo
n de Investigaciones Arqueolo
 gicas Nacionales, Bogota
.
y Sueva. Fundacio
gicas en los abrigos
Correal, G., van der Hammen, T., 1977. Investigaciones arqueolo
rocosos del Tequendama. Biblioteca Banco Popular, Bogot
a.
Erdtman, G., 1969. Handbook of Palinology e Morfology e Taxonomy e Ecology.
Hafner Publishing Company, New York.
Erdtman, G., 1971. Pollen Morphology and Plant Taxonomy: Angiosperms (An
Introduction to Palynology I). Hafner Publishing Company, N. Y.
Espinal, S., 1990. Zonas de vida de Colombia. Departamento de Ciencias de la Tierra.

Universidad Nacional de Colombia, Bogot
a. GAIA. Proyecto Desarrollo Vial Aburra
 n Aburra
 Norte S.A. HATOVIAL. Medelln (Unpublished results).
Norte. Concesio
Faegri, K., Iversen, J., 1975. Textbook of Pollen Analysis. Hafner Press, New York.
n Altoandinos de los Paramos de
Florez, M.T., 1999. Atlas de tolitos de la vegetacio
Belmira y Frontino. Departamento de Antioquia. Editado por la Universidad
Nacional de Colombia, Sede Medelln. Colciencias y Universidad de Antioquia.
todos de estudio palinolo
 gico. Universidad de Antioquia.
Fonnegra, R., 1989. Me
Facultad de Ciencias Exactas y Naturales, Medelln.
 Norte. Concesio
 n Aburr
GAIA, 1999. Proyecto Desarrollo Vial Aburra
a Norte S.A.
HATOVIAL. Medelln (unpublished results).
micas en el suroccidente de
Gnecco, C., Salgado, H., 1989. Adaptaciones precera
Colombia. Boletn Museo del Oro 24, 35e72.
 n temprana de bosques tropicales de montan
~ a. Editorial
Gnecco, C., 2000. Ocupacio
n.
Universidad del Cauca, Popaya
nicos
Gnecco, C., Aceituno, F.J., 2004. Poblamiento temprano y espacios antropoge
rica. Complutum 15, 151e164.
en el norte de Surame
Groot de Mahecha, A.M., 1992. Checua: una secuencia cultural entre 8.500 y 3.000
~ os antes del presente. Fundacio
 n de Investigaciones Arqueolo
gicas Nacioan
 de Bogota
.
nales, Santafe
Herrera, L., Bray, W., Cardale, M., Botero, P., 1992. Nuevas fechas de radiocarbono
mico de la Cordillera Occidental de Colombia. Archeology and
para el precera
environment in Latin America. Institut voor Pre- en Protohistoristche Acheologie Albert Egges van Giffen. Universiteit van Amsterdam, pp. 145e163.
Herrera, L.F., Urrego, L.E., 1996. Atlas de polen de plantas tiles y cultivadas de la
Amazonia colombiana. Estudios en la Amazonia Colombiana, XI. TropenbosColombia, Bogot
a.
~ a, O., Caldero
n, D., Mendoza, L., Seplveda, L., Rodrguez, D.,
Herrera, L., Bray, Pen
Acosta, C., 2011. Informe bimestral de actividades (Noviembre e Diciembre de
gico Aerocafe
. Presentado a la Asociacio
n
2010) Proyecto de Rescate Arqueolo
 (Unpublished results).
del Aeropuerto del Cafe
, Costa Rica.
Holdridge, L.R., 1967. Life Zone Ecology. Tropical Science Center, San Jose
n del ingle
s por Humberto Jime
nez Saa: Ecologa Basada en Zonas de
Traduccio
, Costa Rica: IICA, 1982.
Vida, 1a. ed. San Jose
nchez, J., Buckler, E., Doebley, J., 2002.
Matzuoko, Y., Vigoroux, Y., Goodman, M.M., Sa
Asingle domestication for maize shown by multilocus microsatelite genotyping.
Proc. Nat. Acad. Sci 99 (9), 6080e8084.
Monsalve, C.A., 2000. Catalogo preliminar de tolitos producidos por algunas
plantas asociadas a las actividades humanas en el sureste de Antioquia,
nicas Forestales y del Medio Ambiente 15, 103e116.
Colombia. Cro
Monsalve, C.A., 2008a. Estudio paleoambiental de un perl con 5 muestras en el
 gico La Morena. Municipio de Envigado, Antioquia, Colombia.
sitio arqueolo
n para la Cultura. Seccio
n Archivo Histo
rico. Municipio de
Secretara de Educacio
Envigado (Unpublished results).
 siles contenidos en instrumentos lticos y
Monsalve, C.A., 2008b. Estudio de microfo
gico Reserva Ecolo
gica La Morena.
su signicado ambiental en el sitio arqueolo
n para la Cultura.
Municipio de Envigado, Antioquia. Secretara de Educacio
n Archivo Histo
 rico. Municipio de Envigado (Unpublished results).
Seccio
n morfolo
 gica de frutos y semillas fosilizadas del
Monsalve, C.A., 2009. Identicacio
gico Reserva Ecolo
gica La Morena. Municipio de Envigado,
contexto arqueolo
 n para la Cultura. Seccio
n Archivo Histo
rico.
Antioquia. Secretara de Educacio
Municipio de Envigado (Unpublished results).
Mnera, L.C., Monsalve, O.D., 1996. Arqueologa de rescate. Va Alterna Troncal de
nico-Parma-Ro Campoalegre. INVIAS-INTEOccidente. Sector Puente Dome
GRAL S.A., Medelln (Unpublished results).
Murillo, M.T., Bless, M., 1974. Spores of recent colombian Pteridophyta. I. Trilete
spores. El Cuaternario de Colombia III, 223e269.
Murillo, M.T., Bless, M., 1978. Spores of recent colombian Pteridophyta. II. Monolete
spores. El Cuaternario de Colombia VI, 319e365.
Nelson, M.C., 1991. The study of technological organization. In: Archaeological
Method and Theory, vol. 3. The University of Arizona Press, pp. 57e100.
n arqueolo
 gica y Plan de
Nieto, L.E., Restrepo, J.C., Restrepo, A., 2003. Prospeccio

Manejo en el Area
Fsica del Proyecto Plan Parcial Pajarito. EDU. Universidad de
Antioquia, Medelln (Unpublished results).
Otero de Santos, H., Santos, G., 2012. Din
amica de cambio en las sociedades prectrico.
hisp
anicas de la cuenca baja del Porce. In: Porce III. Proyecto hidroele
Estudios de arqueologa preventiva. Empresas Pblicas de Medelln E.S.P. Universidad de Antioquia, Medelln, pp. 12e229.
nez, J.R., 2009. Nuevas Perspectivas sobre las Culturas Bot
Pag
an-Jime
anicas
Precolombinas de Puerto Rico: Implicaciones del Estudio de Almidones en
micas y de Concha. Cuba Arqueologica 2 (2),
Herramientas Lticas, Cera
7e23.
n-Jime
nez, J.R., 2011. Early phytocultural processes in the Pre-Colonial Antilles.
Paga
A Pan-Caribbean survey for an ongoing starch grain research. In: Hofman, C.L.,
Van Duijvenbode, A. (Eds.), Communities in Contact. Essays in Archaeology,

Please cite this article in press as: Santos Vecino, G., et al., Alteration of tropical forest vegetation from the PleistoceneeHolocene transition and
plant cultivation from the end of early Holocene through middle Holocene in Northwest Colombia, Quaternary International (2014), http://
dx.doi.org/10.1016/j.quaint.2014.09.018

G. Santos Vecino et al. / Quaternary International xxx (2014) 1e15


Ethnohistory, and Ethnography of the Amerindian Circum-Caribbean. Sidestone
Press, Leiden, pp. 87e116.
Pearsall, S.M., Chandler-Ezell, K., Zeidler, J.A., 2004. Maize in ancient Ecuador: results of residue analysis of stone tools from the Real Alto site. Journal of
Archaeological Science 31, 423e442.
Piperno, D.R., 1998. Paleoethnobotany in the Neotropics from Microfossils. New
insights into ancient plant use and agricultural origins in the tropical forest.
Journal of World Prehistory 12 (4), 398e443.
Piperno, D.R., 2011. The origins of plant cultivation and domestication in the New
World tropics. Current Anthropology 52 (S4), 453e470.
Piperno, D.R., Pearsall, D.M., 1998. The Origins of Agriculture in the Lowland Neotropics. Academic Press, San Diego.
Piperno, D.R., Ranere, A.J., Holst, I., Hansell, P., 2000. Starch grains reveal early
root crop horticulture in the Panamanian tropical forest. Nature 407,
894e897.
Piperno, D.R., Ranere, A.J., Holst, I., Iriarte, J., Dickau, R., 2009. Starch grain and
phytoliths evidence for early Ninth millennium B.P. Maize from the Central
Balsas River Valley, Mexico. Proceedings of the National Academy of Sciences
United States of America 106, 5019e5024.
Punt, W., Blackmore, S., Nilson, S., Le Thomas, A., 1994. A Glossary of Pollen and
Spore Terminology. LPP Foundation, Utrecht. Contribution, 1 (71).
n humana en Panama
 a
Ranere, A., y, Cooke, R., 1995. Evidencias de ocupacio
s, Cavelier,
postrimeras del Pleistoceno y a comienzos del Holoceno. In: Ine

rica tropical.
Santiago, Mora (Eds.), Ambito
y ocupaciones tempranas de la Ame
n Erigaie, Sante Fe
Instituto colombiano de Antropologa, Colcultura, Fundacio
de Bogot
a, Colombia, pp. 5e26.
Rodrguez, C., 1991. Patrones de asentamiento de los agricultores prehisp
anicos en
 n, municipio de Chaparral (Tolima). Fundacio
n de investigaciones
El Limo
 gicas Nacionales, Banco de la repblica, Santafe
 de Bogot
arqueolo
a.
Rodrguez, C., 1995. Asentamiento de los bosques subandinos durante el Holoceno

s, Mora, Santiago (Eds.), Ambito
medio. In: Cavalier, Ine
y ocupaciones temrica tropical. Instituto colombiano de Antropologa, Colcultura,
pranas de la Ame
n Erigaie, Santafe
 de Bogota
, pp. 115e123.
Fundacio
Roubick, D.W., Moreno, J.E., 1991. Pollen and Spores of Barro Colorado Island.
Missouri Botanical Garden, Saint Louis, Missouri.
Salgado, H., 1989. Medio ambiente y asentamientos humanos prehisp
anicos en el
Calima Medio. Instituto Vallecaucano de Investigaciones Cientcas, Cali.
mico en el can
~o
n del ro Calima, Cordillera Occidental.
Salgado, H., 1995. El precera

s, Mora, Santiago (Eds.), Ambito
In: Cavalier, Ine
y ocupaciones tempranas de la
rica tropical. Instituto colombiano de Antropologa, Colcultura, Fundacio
n
Ame
 de Bogot
Erigaie, Santafe
a, pp. 91e98.

15

Salomons, J.B., 1989. Paleoecology of Volcanic soils in the Colombian Center Cordillera
(Parque Nacional de los nevados). Studies on Tropandean ecosystems 3, 15e217.
. Prospeccio
n y rescate
Santos, G., 2006. Una tumba de cancel en el Valle de Aburra

gico en el a
rea de la Urbanizacio
 n Alamos
arqueolo
de Escobero. In: Municipio
de Envigado. Envigado (Unpublished results).
Santos, G., 2008. Cazadores-recolectores y Horticultores del Holoceno Temprano y
pez, Carlos, Ospina, Guillermo (Eds.),
Medio en la Cuenca Baja del Porce. In: Lo
 rica. Interacciones Sociedad-Ambiente a Distintas Escalas SocioEcologa Histo
 gica de Pereira, Universidad del Cauca y
Temporales. Universidad Tecnolo
 n.
Sociedad Colombiana de Arqueologa, Pereira, pp. 123e138. CD versio
~ os de ocupaciones humanas en Envigado (Antioquia). El
Santos, G., 2010. 10. 000 an
n Archivo Histo
rico. Secretara de Educacio
n para la
sitio La Morena. Seccio
Cultura, Municipio de Envigado. Envigado.
gicas en El Escobero. Secretara de EduSantos, G., 2011. Investigaciones arqueolo
n para la Cultura y Archivo Histo
rico de Envigado, Municipio de Envigado
cacio
(Antioquia) (Unpublished results).
Stothert, K.E., 1985. Las Vegas culture of coastal ecuador. American Antiquity 50 (3),
613e637.
Stothert, K.E., 1988. La prehistoria temprana de la pennsula de Santa Helena:
nea Antropolo
gica Ecuatoriana. Serie monogra
ca
Cultura Las Vegas. In: Miscela
10Museos del Banco Central del Ecuador, Guayaquil, pp. 203e214.
 n y evaluacio
 n del impacto arqueolo
 gico en el yaciTabares, D., 2003. Prospeccio
n Base de Colombia Mo
 vil, Santa Rosa de Cabal (Unmiento La Playa. Estacio
published results).
 n del Cauca Medio,
Tabares, D., 2004. Arqueologa del mundo Arcaico en la regio
 n Ro Campoalegre). Fundacio
n de Investigaciones
Colombia (Prospeccio
gicas Nacionales. Banco de la Repblica, Bogot
Arqueolo
a (Unpublished results).
n
Tabares, D., 2012. Mundo Arcaico en el Cauca Medio Colombia. Investigacio
gica en la cuenca del ro Campoalegre. Editorial Acade
mica Espan
~ ola,
arqueolo
Madrid.
mico temprano en el Cauca
Tabares, D., Vergara, F., 1996. El Jazmn: Un sitio precera
Medio. Monografa de Grado. Departamento de Antropologa. Universidad de
Antioquia, Medelln.
gico San Bernardo (SRC-SB01)
Tabares, D., Restrepo, J., 2003. Yacimiento arqueolo
n y evaluacio
n del impacto arqueolo
gico). EIA Proyecto Parque
(Prospeccio
Cementerio El Jazmn, Santa Rosa de Cabal, Risaralda, Medelln (Unpublished
results).
cnicas de preparacio
n de muestras sedimentarias
Zucol, F.A., Osterieth, M., 2002. Te
 n de Fitolitos. Ameghiniana Revista de la Asociacio
 n Paleonpara la extraccio
 gica Argentina 39 (3), 379e382.
tolo

Please cite this article in press as: Santos Vecino, G., et al., Alteration of tropical forest vegetation from the PleistoceneeHolocene transition and
plant cultivation from the end of early Holocene through middle Holocene in Northwest Colombia, Quaternary International (2014), http://
dx.doi.org/10.1016/j.quaint.2014.09.018

Quaternary International xxx (2014) 1e12

Contents lists available at ScienceDirect

Quaternary International
journal homepage: www.elsevier.com/locate/quaint

Radiocarbon chronology of terminal Pleistocene to middle Holocene


human occupation in the Middle Cauca Valley, Colombia
s Loaiza b, c, Carlos Lo
 pez d,
Ruth Dickau a, *, Francisco Javier Aceituno b, Nicola
Martha Cano d, Leonor Herrera e, Carlos Restrepo f, Anthony J. Ranere c
a

Department of Archaeology, University of Exeter, Laver Bldg., North Park Rd., Exeter EX4 4QE, UK
Grupo Medioambiente y Sociedad, Laboratorio de Arqueologa, Departamento de Antropologa, Universidad de Antioquia, Calle 67 No 53-108, AA 1226
Medellin, Antioquia, Colombia
c
Department of Anthropology, Temple University, Philadelphia, PA 19122, USA
d
rica y Patrimonio Cultural, Facultad de Ciencias Ambientales, Universidad Tecnolo
gica de Pereira, Pereira, Risaralda,
Laboratorio de Ecologa Histo
Colombia
e
gico Aerocaf
Proyecto de Rescate Arqueolo
e, Calle 8 # 5-04, Palestina, Caldas, Colombia
f
Manzana 47-Casa 14 Villa del Prado, Pereira, Risaralda, Colombia
b

a r t i c l e i n f o

a b s t r a c t

Article history:
Available online xxx

Archaeological research over the past two decades in the Middle Cauca region of central Colombia has
documented numerous preceramic sites dating from the terminal Pleistocene to middle Holocene, along
with substantial artifactual and archaeobotanical evidence for early plant use and food production. We
present a radiocarbon chronology of 26 sites, including dates previously available only in unpublished
reports, and 36 new AMS dates from 11 sites. This chronology solidly establishes the preceramic (before
3600 14C BP) human occupation in the Middle Cauca. The earliest date clearly associated with cultural
evidence of occupation is 10,619 66 14C BP at the site of Cuba. Four sites show occupation before 10,000
14
C BP, but between 10,000 and 9000 14C BP, this number increases to eleven sites. Thereafter, despite
evidence of episodic volcanic activity, there is a relatively constant and continuous sequence of human
occupation in the region, although small localized population movements may have occurred. The
fertility of periodically renewed andisols likely attracted settlement and continued occupation of the
region by people practicing early plant cultivation, based on the archaeobotanical evidence for the early
adoption and use of domesticates.
2014 Elsevier Ltd and INQUA. All rights reserved.

Keywords:
14
C dating
Preceramic occupations
Early Holocene
Middle Holocene
Volcanism
Colombia

1. Introduction
The Middle Cauca archaeological region is situated in the
Northern Andes of Colombia, extending from the Department of
Caldas in the north to the Department of Valle de Cauca in the
south, with the Cauca River as its central axis (Fig. 1). Geographically, it is a very heterogeneous region that includes the lowlands of
the Cauca River valley, the eastern slopes of the Cordillera

* Corresponding author. Present address: HD Analytical Solutions, Inc., 952 Oxford St. W., London, Ontario N6H 1V3, Canada.
E-mail addresses: rdickau@gmail.com (R. Dickau), csfjace@antares.udea.edu.co
 pez),
(F.J. Aceituno), nloaiza@temple.edu (N. Loaiza), cel@utp.edu.co (C. Lo
mcano@utp.edu.co (M. Cano), arqueopalestina@gmail.com (L. Herrera),
arqueologocarlos@hotmail.com (C. Restrepo), ranere@temple.edu (A.J. Ranere).

Occidental, and the western slopes of the Cordillera Central (locally


nico). The Calima Valley is also
known as the Macizo Volca
considered part of the Middle Cauca archaeological region even
though it sits on the western slope of the Cordillera Occidental. In
this paper, we focus on the eastern part of the Middle Cauca, in the
piedmont between the Cauca River and the highlands of the
Cordillera Central, where recent archaeological investigations have
concentrated.
Over the past two decades, numerous preceramic sites dating
from the terminal Pleistocene to the middle Holocene have been
identied through archaeological research in the Middle Cauca
~ o et al., 1997; Rojas and Tabares,
region (e.g. INTEGRAL, 1997; Patin
2000; Tabares and Rojas, 2000; Rodrguez, 2002; Cano, 2004, 2008;
Tabares, 2004; Cardale de Schrimpff et al., 2005; Aceituno and
Loaiza, 2007; Restrepo, 2012, 2013a,b). We use the term

http://dx.doi.org/10.1016/j.quaint.2014.12.025
1040-6182/ 2014 Elsevier Ltd and INQUA. All rights reserved.

Please cite this article in press as: Dickau, R., et al., Radiocarbon chronology of terminal Pleistocene to middle Holocene human occupation in the
Middle Cauca Valley, Colombia, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.12.025

R. Dickau et al. / Quaternary International xxx (2014) 1e12

Fig. 1. Map of the Middle Cauca region showing site locations.

preceramic to refer to all cultural periods prior to the appearance of


the rst ceramics in the Middle Cauca region around 3600 14C BP
~ o, 1996;
(Bray, 1989; Bruhns, 1994; Cano, 1995, 2004; Patin
INTEGRAL, 1997; Restrepo, 2006, 2012; Jaramillo, 2008; Herrera
et al., 2011). Lithic assemblages remained remarkably similar
from ca. 10,000e3600 14C BP (hereafter BP; calibrated dates given
as cal B.C. or cal A.D.) (INTEGRAL, 1997; Aceituno and Loaiza, 2007).
These preceramic assemblages consist of simple stone tools on
akes, aked and/or polished stone hoes (azadas), handstones, and
milling stone bases. Bifacial projectile points and formal tools are
rare (but see Bruhns, 1976; INTEGRAL, 1997; Herrera et al., 2011;
pez and Cano, 2012; Restrepo, 2012). The large number of plant
Lo

processing tools points to an early focus on plants that is not shared


with patterns witnessed in the neighboring Middle Magdalena
 Plateau (Correal and Van der Hammen, 1977;
Valley and the Bogota
 pez, 2008a,b; Lo
 pez and Cano,
Correal, 1986; Nieuwenhuis, 2002; Lo
2012).
The importance of plant resources during the preceramic period
has been conrmed through the recovery of archaeobotanical evidence of economic plants, including several domesticates, by
7000 BP (Aceituno and Loaiza, 2007, 2014b; Aceituno and Lalinde,
2011; Aceituno et al., 2013). The presence of domesticates indicates the early adoption of plant cultivation and horticultural
practices in the region, reecting patterns observed elsewhere in

Please cite this article in press as: Dickau, R., et al., Radiocarbon chronology of terminal Pleistocene to middle Holocene human occupation in the
Middle Cauca Valley, Colombia, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.12.025

R. Dickau et al. / Quaternary International xxx (2014) 1e12

the humid lowland Neotropics (Piperno and Pearsall, 1998;


Piperno, 2011). The Cauca Valley appears to have been an important route of early crop dispersals between Central and South
America during the early to middle Holocene (Aceituno and Loaiza,
2014a).
This paper presents a radiocarbon chronology of the preceramic
period in the Middle Cauca region, from the terminal Pleistocene to
the middle Holocene, based on sites excavated between the Cauca
River and the Cordillera Central. This chronology forms an essential
foundation from which to investigate changes in settlement patterns and subsistence strategies during this critical period,
including the potential impacts of volcanism and other environmental factors. As this is the rst work that comprehensively integrates early sites located in the Middle Cauca, the analysis
presented below is based primarily on the location of sites and
radiocarbon dates.

2. Regional setting and environment


The Middle Cauca region encompasses the terrain drained by
the middle course of the Cauca River from the city of Manizales in
the north to the city of Cali in the south, between the Cordillera
Occidental and Cordillera Central of the Northern Andes. Within
this larger region, the majority of preceramic archaeological sites
have been found in the piedmont zone extending from the Cauca
River oodplain at ~900 m asl (meters above sea level) to the base
of the Central Cordillera at ~2100 m asl, within southern Caldas
Department, eastern Risaralda Department, and northern Quindo
Department (Fig. 1). This area lays in the shadow the highest
mountains in the Northern Andes: the snow-covered volcanic
peaks of Ruz, Huila, Tolima, Quindo, and Santa Isabel of the Central
Cordillera, which rise over 5000 m asl.
The geology of the western slopes of the Central Cordillera is
primarily the result of volcanic activity and subsequent erosion
over the past four million years. Intense tectonic and volcanic
activity during the Paleogene and Neogene melted massive
alpine ice sheets, creating enormous mudows and sedimentation in the Cauca valley (Tilst, 2006), including the large PereiraeArmenia volcanoclastic fan (Guarn et al., 2004; Guarn, 2008).
Around 20,000 years ago, volcanic activity changed to principally
explosive eruptions that produced large amounts of ash. In at
least ve to seven mega-events (averaging one every 3000e4000
years), the whole region was covered by volcanic ash falls of
more than one meter thick (Tilst, 2006). More recent eruptions
during the early and middle Holocene (10,000e3600 BP)
continued to deposit thick ash layers in parts of the region (Herd,
ndez, 1997; Orozco, 2001;
1982; Thouret et al., 1985, 1995; Me
ndez et al., 2002). Because of the amount of volcanic ash
Me
accumulated in the Middle Cauca over the last 10,000 years, early
preceramic deposits are usually deeply buried (~70e160 cm)
(Orozco, 2001; Tilst, 2006). Compacted ash created the presentday rounded relief of the foothills, dissected by small rivers and
streams eroding through the soft sediments. Surface soils are
mainly well-drained, deep andisols that are moderately eroded.
They are acidic (pH 5.2e6.1), but fertile, supporting extensive
modern cultivation of coffee and other crops (Guhl, 1975;
INTEGRAL, 1997; IGAC, 1998). In some regions, the volcanic
sediments have weathered to a red clay, which overlies sandstones, clays, marls, and conglomerates from the Neogene
mudows (Acevedo, 1964). Floodplains are a mix of uvial and
colluvial volcanic sediments.

The region receives 1000e3000 mm of rain annually with two


periods of drier weather in JanuaryeFebruary and JulyeAugust.
Due to prevailing winds, weather patterns, and topography, the
highest rainfall occurs along the piedmont where the most of the
archaeological sites are located (Aceituno and Loaiza, 2007); between 1200 and 1600 m asl, annual rainfall is 2500e3000 mm
(Oster, 1979). Temperatures average 18e24  C with little seasonal
variation. The natural vegetation for the region in the absence of
human interference would be premontane wet forest in the
piedmont and tropical dry (deciduous) forest in the lower elevations along the Cauca River (Holdridge, 1967; Espinal, 1990; IGAC,
1998). The vertical distribution of life zones would have had an
impact on resource accessibility, especially during the climatic
instability of the PleistoceneeHolocene transition (Aceituno and
Loaiza, 2007).
Pollen column samples from the premontane sites of El
Jazmn, Campoalegre, and Guayabito suggest that during the
terminal Pleistocene (>10,000 BP), the climate was cooler and
drier than today, based on high frequencies of Podocarpus,
Junglans nigra, Quercus, Alnus, and Cyathacea (Jaramillo and
Meja, 2000a, 2000b; Aceituno, 2001). The transition to the
early Holocene (10,000e7500 BP) was marked by an increase in
pollen taxa indicative of warmer temperatures and higher humidity. By 9000 BP, climatic conditions were relatively similar to
the present, although altitudinal vegetation zones were
100e200 m lower (Salomons, 1989:166) indicating slightly
cooler conditions. At the super-regional scale, the beginning of
the early Holocene was characterized by an increase in temperature and humidity (Melief, 1985, 1989; Marchant et al.,
2002). Paleoclimatic evidence from the Parque de los Nevados
shows that the start of the middle Holocene (7500e6500 BP)
marked a warmer period with higher temperatures than today
(Melief, 1985; Salomons, 1989; Thouret et al., 1995). This
warmer period, known as the hypsithermal or Holocene Climatic Optimum, has also been documented in palaeoclimatic
records from other parts of the country (Kuhry et al., 1983;
Hooghiemstra and van der Hammen, 1993; Thouret et al.,
1995; Berro et al., 2001; Marchant et al., 2001). Around
6200e6000 BP, changes in vegetation biomes indicates that
there was a short period of cooling known as the Santa Isabel
(Melief, 1985; Salomons, 1989; Thouret et al., 1995), which
continued until 5000 BP (Marchant et al., 2001). Between 4000
and 3000 BP, there was a shift to wetter climatic conditions
which continued until 1000 BP. Between 1000 and 500 BP there
was a drier period in the region (Marchant et al., 2001).

3. Radiocarbon chronology of the Middle Cauca region


We describe 26 sites with dated preceramic deposits from the
Middle Cauca region (Fig. 1). Many of these sites also have signicant
ceramic occupations (post 3600 BP), but these occupations are not
discussed in this paper. In order to examine the preceramic history,
particularly with regards to settlement patterns, landscape interactions, and resource use, we compiled radiocarbon dates from
previously published literature, professional consulting reports, and
unpublished manuscripts (Table 1). Dates from unknown provenience, or those for which we could not track down the original
citation were not included. We also present 36 new AMS dates obtained during our recent excavations in the region (in bold in Table 1).
Calibrations are calculated to two sigma using Calib 7.0 (Struiver et al.,
2005).

Please cite this article in press as: Dickau, R., et al., Radiocarbon chronology of terminal Pleistocene to middle Holocene human occupation in the
Middle Cauca Valley, Colombia, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.12.025

R. Dickau et al. / Quaternary International xxx (2014) 1e12

Table 1
Preceramic radiocarbon dates from the Middle Cauca region.
C Date

Calibrated
date
Yrs BC (2s)

d13C

Method Sample
type

References

AA102509
Beta-285871
Beta-290954
Beta-283582
Beta-317784
Beta-87730
Beta-87729
Beta-104559
Beta-93154
AA98950
AA98951
AA98942
AA98943
Beta-87189
AA98944
AA98945
Beta-87188
Ua-24499

9663
8550
8470
8030
6990
7600
4270
6520
8380
9155
9451
8712
8674
7685
8704
553
9490
8680

10,757e11,220
9446e9657
9439e9533
8635e9126
7795e7937
8202e8535
4581e5038
7272e7570
9136e9532
10,224e10,492
10,520e11,069
9545e9892
9533e9886
8208e8262
9544e9887
515e645
10,505e11,170
9535e9886

24.0
25.5
25.9
26.1
27.5
N/A
N/A
25.0
25.0
26.3
28.8
26.2
26.7
N/A
25.9
25.7
N/A
26.7

AMS
Cnvtl.
AMS
Cnvtl.
Cnvtl.
Cnvtl.
Cnvtl.
AMS
Cnvtl.
AMS
AMS
AMS
AMS
Cnvtl.
AMS
AMS
Cnvtl.
AMS

Herrera et al., 2011


Herrera et al., 2011
Herrera et al., 2011
Herrera et al., 2011
Herrera et al., 2011
INTEGRAL, 1997
INTEGRAL, 1997
Rodrguez, 1997: 94
INTEGRAL, 1997

13 (80e90)
12 (80e85)

Beta-87508
Ua-24494

5825 70
4715 45

6454e6789
5321e5583

N/A
27.4

Cnvtl.
AMS

Charcoal
charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal/
sediment
Charcoal
Charcoal

A1eB1

14 (90e96)

Ua-24495

5625 50

6302e6496

28.5

AMS

Charcoal

N/A
3
3

N/A
C1
Floor

16, Strat V
17 (105e110)
18 (112e117)

Beta-95602
AA98946
Ua-24496

7590 60
7528 51
7080 50

8219e8541
8203e8414
7796e8000

El Jazmn
El Jazmn
El Jazmn

3
N/A
3

D1
N/A
D1

19 (115e120)
21, Strat VI
21 (125e130)

AA98947
Beta-95061
Ua-24497

8660 55
9020 60
10,120 70

Guayabito
Guayabito
La Pochola

N/A
N/A
1 (Pit)

N/A
N/A
1

8, Strat III
Strat V
7 (55e60)

Beta-95063
Beta-95064
Ua-24498

4180 70 4526e4854
7990 100 8589e9125
8095 55 8777e9248

La Pochola
La Pochola

1
1

B2
B1

8 (50e55)
12 (70e75)

AA98943
LTL4221A

5922 51
6743 45

6650e6884
7515e7674

La Pochola

A4/B4

15 (85e90)

LTL4222A

6903 45

7659e7842

La Pochola
La Pochola
La Pochola
La Pochola
San Germ
an

1
1
1
1
1

B1/B2
A3/B3
B4
B3/B4
1

17 (95e100)
18 (100e105)
20 (110e115)
21 (115e120)
9 (75e80)

LTL5436A
LTL4223A
AA98952
LTL4224A
CSIC1987

La Romelia
La Chillona
UTP Bosque
Deportes
~ ita
La Montan
~ ita
La Montan
Nuevo Sol
Cuba
Cuba
Cuba
Cuba
Cuba
Cuba
Cuba
Cuba
Cuba
Cuba
Cuba
Cuba
Cuba
Cuba
*Cuba
UTP Jardn
nico
Bota
La Mikela
La Mikela

2
1
1

B
A
B1

8 (90e100)
7 (80e90)
20 (95e100)

Beta-325215
Beta-325216
AA98954

1
2
I
Exc-1998
20
18
11
19
20
22
19
17
19
U1-1998
10E
20
10E
10
1

A
A
D
several

13 (70e80)
7 (90e100)
(70e80)
(45e55)
11 (50e55)
13 (60e65)
15 (70e75)
16 (75e80)
19 (90e95)
20 (95e100)
21 (100e105)
21 (100e105)
28 (135e140)
27e28 (130e140)
28 (135e140)
31 (152)
33 (160e163)
(164)
12 (110e120)

Beta- 151344
Beta-325214
Beta-306257
Beta-123078
AA102505
AA102502
AA102501
AA102503
AA102497
AA103316
AA102500
AA102499
AA102504
Beta-121972
AA102496
AA102498
AA102510
AA102508
AA98955

4 (35e40)
8 (60e65)

AA98936
AA98937

Site

Block

El Mirador
El Recreo Cancha
El Recreo Cancha
El Recreo Cancha
El Perro
Campoalegre
Campoalegre
Los Arrayanes
El Antojo
Invas 3
Invas 3
La Selva
La Selva
La Selva
La Selva
*La Selva
La Selva
La Selva

Level
(depth cm)

Lab code

Monitoreo 117
C-15
(several)
C-15
C1-21/C2-21
C-15
Feature D-18
Cut 10
Pit
1
1
1
N/A
1
N/A
1
1
1
1
1
1
LS12
N/A
1
1
1
2
LS11
N/A
1
C1 (oor)

C3 (100e110)
(109e147)
(150)
(171e211)
(123e143)
N/A
N/A
(66e77)
17
7 (45e50)
8 (50e55)
7 (55e60)
8 (60e65)
8e9 (60e70)
9 (65e70)
10 (70e75)
10e11 (70e80)
12 (80e85)

*La Selva
El Jazmn

LS8
3

N/A
B1

El Jazmn

El Jazmn
El Jazmn
El Jazmn

E2
0N1W

Quad.

350-352E

B2

14

9047
9312
13,098
13,540
8136

83
60
40
80
30
90
70
90
90
57
58
60
61
110
56
37
110
60

45
55
75
60
65

7630 40
8200 40
4393 44
7300
9230
8740
4220
7007
5780
7014
5844
6990
7466
5863
7001
5911
9730
9826
7032
10619
6460
9284

70
50
50
180
53
49
63
50
57
43
55
53
49
100
63
54
66
51
58

3746 49
4794 45

25.0 Cnvtl.
27.3 AMS
28.3 AMS

Charcoal
Charcoal
Charcoal

9528e9856
26.9 AMS
9919e10,271
25.0 Cnvtl.
11,396e12,027
26.5 AMS

Charcoal
Charcoal
Charcoal

25.0
25.0
27.5

Cnvtl.
Cnvtl.
AMS

25.1 AMS
23.5 AMS
23.8

AMS

Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal

10,159e10,269
26.1 AMS
10,297e10,668
26.8 AMS
15,384e15,980 25.0 AMS
16,092e16,549
22.0 AMS
8790e9089
25.9 Cnvtl.

Charcoal
Charcoal
Charcoal
Charcoal
Charcoal

8375e8536
9027e9277
4852e5269

25.0 AMS
26.6 AMS
26.1 AMS

Charcoal
Charcoal
Charcoal

6355e6021
10,251e10,544
9556e9901
4289e5306
7713e7945
6448e6714
7704e7955
6504e6776
7697e7936
8192e8372
6505e6795
7707e7939
6638e6880
10,746e11,320
11,137e11,395
7738e7962
12,424e12,708
7271e7457
10,270e10,649

28.2
25.7
25.6
25.0
26.0
26.6
26.8
25.9
25.7
25.4
25.2
22.7
26.6
25.0
26.4
24.4
29.7
25.6
25.2

Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Sediment
Charcoal

3930e4247
5331e5605

26.2 AMS
26.8 AMS

AMS
AMS
AMS
Cnvtl.
AMS
AMS
AMS
AMS
AMS
AMS
AMS
AMS
AMS
Cnvtl.
AMS
AMS
AMS
AMS
AMS

INCIVA, 1995e1996

INCIVA,1995e1996
Aceituno y
Loaiza, 2007
INCIVA, 1995e1996
Aceituno and
Loaiza, 2007
Aceituno and
Loaiza, 2007
INTEGRAL, 1997
Aceituno and
Loaiza, 2007
INTEGRAL, 1997
Aceituno and
Loaiza, 2007
INTEGRAL, 1997
INTEGRAL, 1997
Aceituno and
Loaiza, 2007
Aceituno and
Lalinde, 2011
Aceituno and
Lalinde, 2011
Aceituno, 2010
Aceituno, 2010
Aceituno, 2010
Aceituno y
Loaiza, 2007
Restrepo, 2013a
Restrepo, 2013a

CISAN, 2001
Restrepo, 2013a
Restrepo, 2013a
Cano, 2004

Cano, 2004

Charcoal
Charcoal

Please cite this article in press as: Dickau, R., et al., Radiocarbon chronology of terminal Pleistocene to middle Holocene human occupation in the
Middle Cauca Valley, Colombia, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.12.025

R. Dickau et al. / Quaternary International xxx (2014) 1e12


Table 1 (continued )
Level
(depth cm)

Lab code

14

(70e80)
9 (70e80)
14 (90e95)
22 (105e110)
(100e110)
29
7 (75e85)
8 (90e100)
3 (45e50)
8 (90e100)
7 (75e85)
10

Beta-145285
AA98938
AA98939
AA98940
AA103318
Beta-181065
LTL4267A
LTL4845A
AA102607
Beta-325213
Beta-325217
Beta-146612

5850
7208
10,376
2582
3886
4270
9542
9333
5517
10,130
9230
7400

Salento 24

15

Beta-146613

Salento 21

N/A
(190e220)

Site

Block

La Mikela
La Mikela
La Mikela
*La Mikela
*La Mikela
*La Mikela
La Trinidad I
La Trinidad II
El Guatn
El Guatn
Genova
Salento 24

0N1W
E
1N1W
E2
E
0N2W
1
2
1
1
1

Chaguala

Quad.

B
A
A

road cut

Calibrated
date
Yrs BC (2s)

d13C

Method Sample
type

References

6508e6783
7943e8164
11,990e12,525
2499e2772
4160e4420
4654e4917
10,699e11,096
10,298e10,708
6213e6407
11,408e12,023
10,261e10,510
8046e8364

25.0
27.2
26.4
26.1
28.6
26.0
22.9
22.9
29.6
25.9
25.7
N/A

Cnvtl.
AMS
AMS
AMS
AMS
AMS
AMS
AMS
AMS
AMS
AMS
Cnvtl.

Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal
Charcoal

Cano, 2004

9680 100 10,733e11,245 N/A

Cnvtl.

Charcoal

Beta-146609

8430 100 9135e9552

N/A

Cnvtl.

Charcoal

AA103317

7427 40

24.7 AMS

C Date

50
58
70
40
39
40
50
65
49
50
40
70

8178e8343

Restrepo, 2013a
Restrepo, 2013a
Restrepo, 2013a
Restrepo, 2013a
Rojas and
Tabares, 2000
Rojas and
Tabares, 2000
Rojas and
Tabares, 2000

Charcoal

Notes: Samples in bold are new unpublished dates from the Middle Cauca Archaeology Project. Samples marked with an asterisk (*) are rejected dates (see text).
Cnvtl. Conventional date, AMS Accelerated Mass Spectrometry date. Calibrations calculated to 2s using Calib 7.0 (Struiver et al., 2005).

3.1. El Mirador (AP 11) (1575 m asl, >2000 m2)


El Mirador is located in the Municipality of Palestina, Department of Caldas, on the northern-most summit in a range of hills
 River valley. To the north and south,
rising above the Chinchina
these hills are bordered by tributaries of the Sirena creek. Several
stone hoes (azadas) were found during surface collecting. During
excavation, an anthrosol was identied 2.5 m below surface, containing charcoal, broken (split) stones, a grinding stone, and unused
akes. A chert projectile point was recovered from the erosional
surface of the anthrosol during post-construction survey (Herrera
et al., 2009, 2011). An AMS assay on charcoal dates the anthrosol
to 9663 83 BP.
3.2. El Recreo Cancha (AP 39) (1585 m asl, 2500 m2)
This site is located on a attened hilltop in the same range of
hills as the El Mirador site, approximately 1 km to the south. Preceramic deposits at the site are represented by a sequence of
anthrosols, beginning >1.5 m deep. Horizon Ab6, associated with a
date of 8470 40 BP, contained small amounts of lithic material
and organic matter. Artifact density increases in Horizon Ab5,
associated with a date of 8550 60 BP, and then decreases in Horizon Ab4, dating to 8030 80 BP. The majority of lithics recovered
from these levels are split stones, but there are also fragments of
tools, debitage, hoes, milling stone bases, cobble handstones, and a
fragment of a quartz projectile point (Herrera et al., 2011). Among
the macrobotanical remains identied were seeds from a type of
bana (Annona sp.), blackberry or mora (Rubus sp.),
soursop or guana
legume (Fabaceae), avocado (Persea americana), palm (Arecaceae)
wood, and a small unidentied tuber or root (Morcote et al., 2010).
3.3. El Perro (AP 107) (1570 m asl, 15 m2)
El Perro is located on a ridge that descends east from El Recreo
Cancha site located some 150 m upslope. The ridge has been greatly
impacted by recent tectonic activity, along with modern agricultural and construction activities, so that the remnants of an identied anthrosol are limited to a small undisturbed area. The original
extent of this anthrosol, buried under 1 m of volcanic ash, may have
been much greater, given that the terrace is ~2700 m2. Within the

remaining anthrosol, excavations uncovered a pit feature lled with


charcoal, a sample of which dated to 6990 30 BP. Directly above
the pit was a concentration of coarse broken or crushed stone
(sandstone and quartz), and a hoe and lithic debitage were recovered nearby (Herrera and Moreno, 2011:10e15).
3.4. Campoalegre (1400 m asl)
The site of Campoalegre is located in the municipality of
, Department of Caldas, on a mid-level terrace east of the
Chinchina
Campoalegre River. The preceramic deposits are represented by
Strata IV, dated to 4270 70 BP, and Strata IVa, dated to
7600 90 BP. In both layers, the lithic assemblage is composed of
handstones, hoes, and ake tools (INTEGRAL, 1997:30e33).
3.5. Los Arrayanes (2400 m asl)
Los Arrayanes is the highest altitude site discussed in this paper,
located in the municipality of Villamara, Department of Caldas, on
a colluvial terrace of the Cordillera Central. Initial occupation of the
site is represented by preceramic lithic assemblage recovered
121e189 cm b.s., which is similar to other sites in the region and
includes handstones, edge-ground cobbles, anvils, grinding bases,
akes, scrapers, and a fragment of hoe (Rodrguez, 1997:116e119).
After a period of abandonment (112e121 cm b.s.), the site was
reoccupied, based on material recovered between 112 and 60 cm
b.s. A date of 6520 90 BP was obtained from Horizon Ab3
(90e60 cm) associated with the end of this occupation. A third
preceramic occupation was identied in the upper levels of the site
(Rodrguez, 1997).
3.6. El Antojo (1440 m asl, 1200 m2)
El Antojo, located on top of a hill on the west side of the Campoalegre River, includes an unprecedented quartz workshop
(INTEGRAL, 1997). This site appears to be the source of the few
n
quartz artifacts found at nearby sites like El Jazmn and San Germa
(Aceituno and Loaiza, 2007). Along with signicant amounts of
quartz debitage, a quartz bifacially-aked preform was recovered,
along with a notched hoe similar to those found at other sites in the
region (Tabares, 2004). A date of 8380 90 BP on charcoal was

Please cite this article in press as: Dickau, R., et al., Radiocarbon chronology of terminal Pleistocene to middle Holocene human occupation in the
Middle Cauca Valley, Colombia, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.12.025

R. Dickau et al. / Quaternary International xxx (2014) 1e12

obtained from Level 17 (INTEGRAL, 1997), but initial occupation of


the site occurred earlier, based on cultural material recovered in
levels below this date. Despite the presence of the hoe, this site
appears to have primarily been a workshop, rather than a domestic
campsite like the majority of other sites.
3.7. Invas 3 (1582 m asl, 650 m2)
This site is located on top of a hill on the west side of the
Campoalegre River, approximately 300 m west of the site of El
Antojo. Much of the original extent of the site has been disturbed by
modern construction, but within a small undisturbed area, excavation revealed a stratied preceramic deposit 30e60 cm below
surface, containing handstones, two milling stone bases, one hoe,
irregular cores, and akes, including quartz akes. Two AMS dates
from Invas 3 show an intensive early Holocene occupation, between 9451 58 BP (Level 8) and 9155 57 BP (Level 7).
3.8. La Selva (1600 m asl, 10,000 m2)
This site is located on a high terrace of colluvial deposits on the
eastern slope of the Mil Ochenta ridge, just north of the town of
Marsella in the Department of Risaralda, with a broad view of the
Cauca River valley (Aceituno and Loaiza, 2007). Excavations
recovered buried preceramic levels containing handstones, milling
stone bases, hoes, and debitage, primarily of basalt with some
andesite. Despite its location near the El Antojo quartz workshop,
no quartz artifacts were recovered.
Initial dating of the site in 1995 indicated occupation between
9500 and 7700 BP (INCIVA, 1995e1996; Aceituno and Loaiza,
2007). However, there was evidence of stratigraphic disturbance
in the lowest levels, where a date of 5825 70 BP (Level 13) was
obtained below dates of 8680 60 BP (Level 12) and 9490 110 BP
(Level 10e11). This earlier date was therefore rejected. Level 8e9
dated to 7685 110 BP. We obtained four new AMS dates in 2013 on
charcoal. Three of these cluster tightly around 8700 BP (Level 7,
8712 60 BP; Level 8, 8674 61 BP; Level 9, 8704 56 BP) and are
generally within the range expected based on stratigraphy. However, a new charcoal sample from Level 10 yielded a very late date of
553 37 BP and again indicates some sort of disturbance or
modern intrusion, and is therefore rejected. Despite the apparent
stratigraphic disturbance, all dates taken together show that the
most intense period of occupation of the site occurred between
8700 and 7700 BP with a likely initial occupation by 9500 BP.
Recent starch analysis of tool residues show the use of Dioscorea sp.,
Phaseolus sp., and cf. Manihot sp. by 8700 BP (Aceituno and Loaiza,
2014a), supporting Aceituno and Loaiza's (2007) original interpretation of the site as a residential campsite.
3.9. El Jazmn (1650 m asl, 2000 m2)
The site is located on a hill just east of the San Eugenio River
(INTEGRAL, 1997; Aceituno and Loaiza, 2007) approximately 4 km
north of the town of Santa Rosa de Cabal, Department of Risaralda.
Excavations uncovered a stratied preceramic deposit 70e140 cm
below surface, which contained handstones, grinding stone bases,
notched hoes, ake tools, and debitage. This deposit dated from
10,120 70 BP to 4715 45 BP, based on six original radiocarbon
dates (Rodrguez, 2002; Aceituno and Loaiza, 2007). Two new AMS
dates obtained for El Jazmn e 8660 55 BP from Level 19 and
7528 51 BP from Level 17 e t reasonably well with the original
chronology. There may have been some slight mixing within Levels
16e18, but the dates from these levels are all within 500 years of
each other. In general, the chronological sequence is consistent
with the site stratigraphy. The dates and associated cultural

material indicate several major phases of occupation: initial colonization from 10,120 70 BP to 8660 55 BP, an intensive period of
occupation from 7590 60 to 7080 50 BP based on the density of
cultural material, and a later phase from 5625 50 to 4715 45 BP.
Pollen analysis at the site shows forest disturbance by 7000 BP, and
the presence of Zea mays and Xanthosoma sp. between 7000 and
5000 BP. Starch grains from tools conrm the availability of maize
by 7000 BP, and show that Manihot sp., Dioscorea sp., and Phaseolus
sp. were used even earlier, by 7600 BP (Aceituno and Loaiza, 2014a).
3.10. Guayabito (1623 m asl, 820 m2)
This site is situated on a small colluvial terrace overlooking the
San Eugenio River, approximately 200 m south of El Jazmn. The
preceramic deposits were divided into three Strata: III, IV, and V.
Stratum III, dating to 4180 80 BP, shows the most intense occupation of the site, with an artifact assemblage of grinding stones,
hoes, and akes. Pollen of Zea and Manihot were identied in this
level (Aceituno, 2002). Stratum V produced a charcoal date of
7990 100 BP but this stratum and stratum VI above it contained
no lithic artifacts (INTEGRAL, 1997).
3.11. La Pochola (1677 m asl, 1500 m2)
La Pochola is located on a rounded terrace of uvial-volcanic
origin west of the San Eugenio River, 1.4 km southwest of El Jazmn and Guayabito. The original excavations of the site recovered a
preceramic lithic assemblage composed of handstones, grinding
bases, aked tools and debitage from levels (18e7) dating between
9312 55 BP and 6743 45 BP (Aceituno and Loaiza, 2007;
Aceituno, 2010; Aceituno and Lalinde, 2011). A date of
13,540 60 BP was obtained from Level 21, but was not associated
with any cultural material and is therefore presumably before
initial occupation of the site (Aceituno, 2010). A date from Level 7 of
8095 55 BP was out of stratigraphic sequence; however, it was
associated with a pit feature and therefore represents a disturbed
context.
Two new dates were obtained from La Pochola to supplement
the previous ve from the original excavation. The rst date of
13,098 75 BP (Level 20) is consistent with the previously early
date obtained from Level 21 just below, and likewise was not
associated with cultural material. The second new date,
5922 51 BP (Level 8), extends the known preceramic occupation
of the site and is consistent with late preceramic dates from other
sites in the region, such as El Jazmn, La Mikela, and Cuba. The
distribution of dates and cultural material suggests that there were
at least two major phases of occupation: initial colonization of the
site 9312 55 BP to 9047 45 BP, and a second, more intensive
occupation (based on artifact density) 8095 55 BP to 5922 51 BP
(Aceituno and Loaiza, 2007). This pattern of settlement is similar to
that of nearby El Jazmn (Aceituno and Loaiza, 2007). Starch analysis of preceramic tools at the site shows the presence of Phaseolus
sp., Dioscorea sp., and Z. mays by 6700 BP (Aceituno and Lalinde,
2011; Aceituno and Loaiza, 2014a).
n (1649 m asl, 800 m2)
3.12. San Germa
n is located on a hilltop of uvial-volcanic
The site of San Germa
origin west of the San Eugenio River, less than 1 km south of La
Pochola. Excavation yielded a localized concentration of artifacts,
including grinding tools, debitage, and a few aked tools, from a
25 cm thick preceramic deposit. A single radiocarbon date of
8136 65 BP on charcoal was obtained from this deposit (Aceituno
and Loaiza, 2007, 2014a).

Please cite this article in press as: Dickau, R., et al., Radiocarbon chronology of terminal Pleistocene to middle Holocene human occupation in the
Middle Cauca Valley, Colombia, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.12.025

R. Dickau et al. / Quaternary International xxx (2014) 1e12

3.13. La Romelia (1480 m.s.n.m. 1500 m2)


La Romelia is located in the municipality of Dosquebradas, on
the slope of a terrace located between two attened hilltops east of
La Chillona Creek. During excavations, the preceramic occupation of
the site was identied 50e130 cm below surface, containing lithic
material and charcoal. In Horizon Ap (Levels 7e9), a paleosol was
detected which yielded a large amount of lithic tools, debitage, and
charcoal. Within this paleosol, a charcoal sample from Level 8
(90e100 cm) was dated to 7630 40 BP (Restrepo, 2008, 2013a).
3.14. La Chillona (1550 m asl, 350 m2)
La Chillona is located on the northern edge of a terrace of volcanic origin overlooking La Chillona Creek, approximately 500 m
east of La Romelia. Excavation recovered a signicant assemblage of
lithic material 5e70 cm deep, including tabular milling stone bases,
debitage, and rounded stones. A carbon sample from just below the
preceramic deposit at 80 cm dated to 8200 40 BP, representing
the initial occupation of the site (Restrepo, 2008, 2013a).
3.15. UTP Bosque Deportes (66PER016) (1470 m asl, ~3000 m2)
This site, located in a forested area north of the main sports eld
 gica de Pereira (UTP), is situated on a
of the Universidad Tecnolo
northern remnant of the PereiraeArmenia geologic fan. The site
was discovered during an archaeological survey within the campus
when cultural materials, including edge-ground cobbles indicative
of preceramic occupation, were recovered eroding out of a cut bank
and in modern pit disturbances. Excavations of a small test unit
revealed preceramic lithics and charcoal in stratigraphic levels
between 55 and 120 cm b.s. A carbon sample associated with a
cobble tool at 96 cm dated to 4393 44 BP, showing the site was
occupied near the end of the preceramic period.
~ ita (1230 m asl, 120 m2)
3.16. La Montan
~ ita is located in the Condina District of the city of
La Montan
Pereira, in the barrio of Cuba, on the shoulder of a terrace near the
top of a large hill bounded to the south by the Naranjal stream.
Modern construction activity uncovered an accumulation of lithic
fragments, remnants of a preceramic campsite. Excavations recovered additional lithics, as well as an avocado (P. americana) pit. A
charcoal sample from 90 to 100 cm b.s. showed that the site was
occupied by at least 9230 50 BP (Restrepo, 2013a). A second date
of 7300 70 BP from 70 to 80 cm b.s. in a nearby unit shows
continued human activity at the site (CISAN, 2001).
3.17. Nuevo Sol (1291 m asl, 900 m2)
This site is located on the Nuevo Sol coffee farm, east of the
 highway. It is situated on the edge of a hilltop terrace
Pereira-Alcala
bounded by an intermittent stream. Removal of 50 cm of topsoil by
construction equipment revealed a lithic scatter in the southeast
part of the terrace. Subsequent archaeological excavation uncovered preceramic lithic material and charcoal below this exposed
surface. A charcoal sample from Level 7 (70e80 cm below the
exposed surface) dated to 8740 50 BP (Restrepo, 2013a).
3.18. Cuba (66PER001) (1280 m asl, 10,000 m2)
Cuba is located on a low terrace of eroded volcanic ash deposits,
at the juncture of two small streams, Cundina and El Oso, within the
city limits of Pereira. It was rst identied and excavated in
1998e2001 by Cano (Cano, 1998, 2001, 2004, 2008; Cano et al.,

2001, 2013), with new excavations conducted in 2012. Preceramic


deposits occur 50e170 cm below surface, with a lithic assemblage
that includes handstones, milling stone bases, hoes, ake tools,
irregular cores, bipolar cores, debitage, hammerstones, and a
number of pebble and cobble manuports. An anthrosol was clearly
visible at the base of the main excavation block around 160e165 cm
below surface. A date of 9730 100 BP in 2001 from the original
excavations and two new dates of 10,619 66 and 9826 63 BP
associated with this anthrosol conrm the early occupation of the
site. This anthrosol appears to be very localized, as no dates of
similar antiquity were found in any of the other test pits excavated
several meters away. A fourth date of 6460 51 BP at 164 cm b.s.
within the anthrosol is too young, likely due to the fact that it was a
date on sediment rather than charcoal and is therefore rejected.
The next occupation of the site occurred between 7000 and
5780 BP based on a series of nine new dates, and was spatially more
widespread. However, several of these dates are out of sequence, or
are too young, suggesting they may have been affected by disturbance or bioturbation. During this second occupation, the tight
clustering of dates across signicant depths (up to 100 cm) suggests
that either deposition events (e.g. ashfall or uvial) occurred
regularly and frequently, or this zone experienced signicant
amounts of post-depositional mixing. A date of 4220 180 BP
marks the end of the preceramic period at the site.
nico (66PER090) (1460 m asl, ~10,000 m2)
3.19. UTP Jardn Bota
Like the UTP Bosque Deportes site to the north, this site is
located on a remnant of the PereiraeArmenia volcanoclastic fan
(Guarn et al., 2004), within the Botanical Garden of the UTP
campus. Archaeological materials including handstones, cores,
akes, and ceramics were found eroding out of a cut bank during
survey, prompting the excavation of a 1  2 m unit to a depth of
120 cm b.s. (Cano et al., 2013). A preceramic occupation with lithic
material and charcoal was identied 70e130 cm b.s. A sample of
carbon from the B Horizon (110e120 cm b.s.), associated with lithic
artifacts and re-cracked rock, was AMS dated to 9284 58 BP.
3.20. La Mikela (66PER007) (1386 m asl, 9000 m2)
La Mikela is located on a terrace composed of volcanic ash deposits, 25 m above the Consota River, 300 m west of a salt spring,
southeast of the city of Pereira (Cano, 2001, 2004, 2008; Cano et al.,
2001, 2013). The preceramic levels occur 60e120 cm b.s., with an
artifact assemblage that includes handstones, milling stone bases,
hoes, irregular cores, bipolar cores, hammerstones, used akes, and
cobble manuports. Three dates from below Level 14, including two
new AMS dates, are too young and are considered intrusive due to
bioturbation. Part of the site showed signicant amounts of bioturbation, mostly from insects and roots. However, ve dates on
charcoal from Level 14 (90e95 cm) to Level 4 (35e40 cm), including
four new AMS dates, are in stratigraphic order, and show that the
preceramic occupation of the site occurred from 10,376 70 BP to
3746 49 BP.
3.21. La Trinidad I and II (1417 m asl, 450 m2)
These sites, located near Barrio Montelbano just south of the
city of Pereira, are located on a large elongated terrace with steep
slopes, bounded on one side by the headwaters of an intermittent
stream. At La Trinidad I on the northern side of the terrace, cultural
material including a hoe, akes, and re-cracked rocks were
recovered 50e160 cm. A charcoal sample from Level 7 (75e85 cm)
dated to 9542 50 BP (Restrepo, 2010, 2013a). At La Trinidad II,
several meters away on the south side of the terrace, a date of

Please cite this article in press as: Dickau, R., et al., Radiocarbon chronology of terminal Pleistocene to middle Holocene human occupation in the
Middle Cauca Valley, Colombia, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.12.025

R. Dickau et al. / Quaternary International xxx (2014) 1e12


8000

7000

6000

8000

7000

6000

5000

4000

3000

5000

4000

3000

Mikela, AA98936
Guayabito, Beta-95063
Cuba, Beta-123078
Campoalegre, Beta-87729
Bosque Deportes, AA98954
Jazmn, Ua24494
Mikela, AA98937
Guatn, AA102607
Jazmn, Ua-24495
Cuba, AA102502
Selva, Beta-87508
Cuba, AA102503

Site, Lab No.

Mikela, Beta-145285
Cuba, AA102500
Cuba, AA102504
Pochola, AA98943
Arrayanes, Beta-104559
Pochola, LTL4221A
Pochola, LTL4222A
Cuba, AA102497
Perro, Beta-317784
Cuba, AA102499
Cuba, AA102505
Cuba, AA102501
Cuba, AA102498
Jazmn, Ua-24496
Mikela, AA98938
Montaita, N/A
Salento 24, Beta-146612
Chaguala, AA103317
Cuba, AA103316
Jazmn, AA98946
Jazmn, Beta-95602
Campoalegre, Beta-87730
Romelia, Beta-325215
Selva, Beta-87189

Years Uncal BP

11000

10000

9000

11000

10000

9000

8000

7000

8000

7000

Guayabito, Beta-95064
Recreo Cancha, Beta-283582
Pochola, Ua-24498
San Germn, CSIC1987
Chillona, Beta-325216
Antojo, Beta-93154
Salento 21, Beta-146609
Recreo Cancha, Beta-290954
Recreo Cancha, Beta-285871
Jazmn, AA98947
Selva, AA98943
Selva, Ua-24499
Selva, AA98944

Site, Lab No.

Selva, AA98942
Nuevo Sol, Beta-306257
Jazmn, Beta-95061
Pochola, LTL5436A
Invias 3, AA98950
Genova, Beta-325217
Montaita, Beta-325214
Jardin Botanico, AA98955
Pochola, LTL4223A
Trinidad II, LTL4845A
Invias 3, AA98951
Selva, Beta-87188
Trinidad I, LTL4267A
Mirador, AA102509
Salento 24, Beta-146613
Cuba, Beta-121972
Cuba, AA102496
Jazmn, Ua-24497
Guatn, Beta-325213
Mikela, AA98939
Cuba, AA102510

Years Uncal BP
Fig. 2. a: Graph of the uncalibrated preceramic radiocarbon dates from archaeological sites and volcanic activity in the Middle Cauca region, 7800e3000 BP. b: Graph of the
uncalibrated preceramic radiocarbon dates from archaeological sites and volcanic activity in the Middle Cauca region, 11,000e7800 BP. Uncalibrated archaeological dates with their
error ranges (horizontal black lines) are arranged in chronological order and identied by their site and radiocarbon laboratory code. Vertical grey lines represent dated volcanic
ndez (1997), Me
ndez et al. (2002),
events and grey shading represents error ranges of these events, or broad ranges of activity. Volcanic activity summarized from Herd (1982), Me
Orozco (2001), Thouret et al. (1985, 1995), and the Smithsonian Inst. (2013). See Table 2.

Please cite this article in press as: Dickau, R., et al., Radiocarbon chronology of terminal Pleistocene to middle Holocene human occupation in the
Middle Cauca Valley, Colombia, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.12.025

R. Dickau et al. / Quaternary International xxx (2014) 1e12

9333 65 BP was obtained on a charcoal sample from 90 to 100 cm


b.s., associated with lithic material (Restrepo, 2010, 2013a).
3.22. El Guatn (1440 m asl, 450 m2)
This site is located southeast of Pereira, in the Vereda Guayabal
District, on the side of a small terrace on the east bank of the
Condina creek. The elongated terrace is bordered to the north and
south by patches of forest, steep slopes, and small waterways that
converge in front of the terrace and ow into the Condina creek. As
such, the campsite was strategically situated for collecting natural
resources (Restrepo, 2013a). Excavations in the central part of the
terrace recovered cultural material between 45 and 100 cm b.s. In
Level 8 (90e100 cm b.s.), at the bottom of the preceramic occupation ending at Horizon Ab, a sample of charcoal dated to
10,130 50 BP (Restrepo, 2013a). A new AMS date on a sample of
charcoal from Level 3 (45e50 cm b.s.) marks the end of the preceramic occupation of the site at 5517 49 BP.
3.23. G
enova (1520 m asl, 975 m2)
This temporary campsite is located west of the Condina creek on
nova Ranch, along the Variante Sur Road south of Pereira, in
the Ge
the Vereda Montelargo District. Like La Trinidad, it is situated on
the south side of a large elongated terrace with steep slopes to the
east and west. Excavations on the eastern ank of the terrace
recovered lithic material and re-cracked rock (Restrepo, 2006). A
charcoal sample from Level 7 (70e80 cm) dates the occupation to
9230 40 BP.
3.24. Salento 24 (2070 m asl)
This site is situated on the top of an eroded knoll along the
highway that runs between Pereira and the town of Salento, in the
municipality of Salento, Quindo. A carbon sample from Horizon
ABb2 (Level 15) dated to 9680 100 BP represents the initial settlement of the site, associated with unused akes and microdebitage. A second occupation level above this in Horizon Ab2
contained lithic akes and a hoe. A third preceramic occupation
level in Horizon ABb1 (Level 10) with akes, micro-debitage, a
scraper, rounded cobles, and re-cracked rock was dated to
7400 70 BP (Rojas and Tabares, 2000).
3.25. Salento 21 (2100 m asl)
Salento 21 is located approximately 1.6 km from Salento 24 near
the same highway, on the top of a hill near the community of San
Antonio. The initial preceramic occupation of the site, dated
8430 100 BP, was associated with domestic refuse, including a
hearth, macrobotanical remains (seeds and a type of nut), burnt
rocks, lithic debitage and used akes (Rojas and Tabares, 2000).
3.26. Chaguala (1532 m asl, 1500 m2)
This site is exposed in a road-cut of the Variante Calarc
a-Pereira
road, west of Armenia, Quindo, at the proximal end of the PereiraeArmenia volcanoclastic fan (Guarn et al., 2004). The exposed
prole shows various tephra layers of varying thickness and
composition, representing recurrent episodes of volcanic ash fall,
accumulation, and preservation, with some strata more than 1 m
thick. The site has not been systematically excavated, but lithic
artifacts characteristic of the Middle Cauca preceramic period,
including a bifacially aked chopper/hoe, were found eroding out
of an organically rich layer 230 cm b.s., 30 cm below a coarse sand
layer believed to be lapilli associated with the eruption of Cerro

Machn (27 km away) in 3600 BP (Cano et al., 2013). However, a


carbon sample from the buried organic layer dated to 7427 40 BP,
suggesting signicant erosion of sediments between the deposition
of the organic layer and the Cerro Machn eruption.
4. Discussion
4.1. New dates
Through our recent archaeological eld work in the region, we
obtained 36 new AMS dates from 11 sites (in bold in Table 1). These
additional dates help round out the occupational history or extend
the known preceramic use of particular sites. Moreover, they help
anchor the settlement chronology and establish a baseline for
assessing cultural patterns and changes over time, such as the
adoption of food production and its social and environmental
consequences.
Most of the new AMS dates t well within existing site chronologies with some exceptions, likely the result of bioturbation and
stratigraphic mixing. There was signicant evidence of bioturbation
at La Mikela and Cuba (primarily caused by insects, worms, and
roots), visible up to 1 m below surface. Despite attempts to avoid
sampling in disturbed areas, these sites produced some out-ofsequence dates that were rejected as intrusive (see Sections 3.18
and 3.20, and Table 1). Other sites, such as El Jazmn and La
Pochola, showed some small-scale disturbances in the upper levels
(mainly worms and roots), but these did not affect the preceramic
chronological sequence.
In most cases, wood charcoal was the material sampled for
dating. Because of the humid environment of the region, it is unlikely that old wood effect was an issue. Wood and organic material
rapidly decays in this region unless preserved through carbonization. Likewise, marine or freshwater reservoir effects are not issues
with our samples.
4.2. Chronological distribution of dates
Based on an examination of all accepted radiocarbon dates, the
Middle Cauca region was inhabited throughout the entire preceramic period, from 10,600 to 3600 BP, with little evidence of any
region-wide abandonment (Fig. 2). The earliest date clearly associated with cultural material is 10,619 66 BP at the site of Cuba.
Three other sites show evidence of human activity earlier than
10,000 BP: El Jazmn, La Mikela, and El Guatn. Between 10,000 and
9000 BP, the number of sites with evidence of human occupation
nearly triples from four to eleven, with the addition of El Mirador,
~ ita, UTP Jardn Bota
nico, La
La Selva, Invas 3, La Pochola, La Montan
nova, and Salento 24.
Trinidad I and II, Ge
Evaluating region-wide settlement patterns is challenging
despite the number of sites and dates compiled. For many of the
archaeological sites discussed, only one or two dates were obtained, and these were often focused on establishing earliest
occupation of the site or time(s) of the most intense occupation. For
this reason, apparent gaps in the regional sequence, particularly in
the later part of the preceramic period, are more likely the result of
sampling bias than actual absence of human activity. However,
several sites were more extensively sampled and provide a
sequence of dates through their entire preceramic occupation,
particularly La Selva, El Jazmn, La Pochola, Cuba, and La Mikela.
Examination of the chronologies of these sites allows some limited
interpretations of local settlement patterns.
As described in Section 2, the Middle Cauca region is situated
within a highly active tectonic area, and the majority of the
archaeological sites lie within the shadow of a major volcanic
massif with nine active or dormant volcanoes (Cerro Bravo, Ruz,

Please cite this article in press as: Dickau, R., et al., Radiocarbon chronology of terminal Pleistocene to middle Holocene human occupation in the
Middle Cauca Valley, Colombia, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.12.025

10

R. Dickau et al. / Quaternary International xxx (2014) 1e12

Cisne, Huile, Santa Isabel, Santa Rosa, Quindo, Tolima, and Machn).
Geomorphological study of the landscape and archaeological excavations through thick ash and mudow layers clearly illustrate
the major impacts that volcanic activity has had on site formation
processes (INCIVA, 1995; Cano et al., 2001, 2013; Restrepo, 2006,
2013b; Aceituno and Loaiza, 2007; Cano, 2008). This raises the
question of what sort of impact this history of volcanic events had
on the past human occupation of the region.
Although the occupation of the region spans the entire preceramic, there are some small gaps within the overall regional
radiocarbon chronology. The most signicant of these (>500
radiocarbon years) are 6460e5922 BP (508 radiocarbon years) and
5517e4794 BP (723 radiocarbon years) (Fig. 2). However, as
mentioned, these gaps are most likely the result of sampling biases
rather than indicative of people abandoning the Middle Cauca region. These gaps do not coincide with known volcanic activity,
documented through geomineralogical analyses and tephndez et al., 2002) (Table 2). In fact, when dates of
rachronology (Me
major volcanic eruptions and other volcanic activity are compared
with dates of human occupation, these often coincide (Fig. 2);
people continued living and working in this landscape, despite the
eposidic eruptions, ash falls, and lava ows.
Table 2
History of volcanic activity in the Middle Cauca region 11,500e3000 BP.
Uncal Yrs BP

Volcanic event

Reference

3050 200
3100
3310 150
3600e4700

Cerro Bravo eruption


Quindo eruption
Cerro Bravo eruption
Eruption of lapilli from
Tolima, Quindo, Cerro Bravo,
Santa Isabel
Machn eruptions
Machn eruptions
Block lava from Tolima,
Quindo, Cerro Bravo,
Santa Isabel
Block lava from Santa Isabel
Cerro Bravo eruption
Block lava from Santa Isabel
Block lava from Santa Isabel
Machn eruption
Cerro Bravo eruption
Ruz eruption
Machn eruption
Ash clouds and pumice ows
from Tolima, eruption of Quindo

Smithsonian Inst., 2013


Orozco, 2001
Smithsonian Inst., 2013
Orozco, 2001

4300e4400
4600e5100
5400e7200

5490 475
6250 110
6759 180
7435 100
8450 95
8630 50
8590 115
9740 95
10,000e11,500

ndez et al., 2002


Me
ndez et al., 2002
Me
Orozco, 2001

ndez, 1997
Me
Thouret et al., 1985
ndez, 1997
Me
ndez, 1997
Me
ndez et al., 2002
Me
Thouret et al., 1985
Herd, 1982
ndez et al., 2002
Me
Orozco, 2001

The beginning of the early Holocene (ca. 11,500e10,000 BP) was


marked by signicant eruptions in the region, including Machn ca.
9740 95 BP (Thouret and Van der Hammen, 1983:275; Orozco,
ndez et al., 2002). From ca. 9000e7500 BP, there were
2001; Me
explosive ash and pumice emissions from several volcanoes
(Salomons, 1989:33). Thereafter, volcanic activity shifted to large
lava ows from Santa Isabel and other volcanos, from ca.
ndez,
7200e5400 BP (Thouret and Van der Hammen, 1983:275; Me
1997). After 5100 BP, activity was primarily eruptions from Machn
in the southern part of the massif, and the deposition of lapilli from
eruptions of Tolima, Quindo, Cerro Bravo, Santa Isabel (Orozco,
ndez et al., 2002). The presence of dated evidence of
2001; Me
cultural activity despite these frequent and ongoing episodes suggests that ancient peoples found ways of adapting to the volcanic
conditions throughout the region.
This can be observed at the site level as well. At La Pochola, the
sequence of volcanic ash layers and cultural material indicates
continued occupation throughout major volcanic events. Evidence
of signicant ash falls, documented by an elevated percentage of

phenocrysts in the soil, coincide with those levels (12e16) which


contain the highest density of archaeological materials, interpreted
as the most intensive occupation of the site (Aceituno and Loaiza,
2007). At El Jazmn, Aceituno and Loaiza (2007:43; 2014a)
observed a similar pattern in Levels 14e21. Mineralogical analysis
of the site sediments indicates that there was relatively low volcanic activity during the initial occupation phase, but increased
volcanic activity between 7200 and 5400 BP (Orozco, 2001).
This does not mean that human populations did not suffer from
the effects of volcanic eruptions. It is likely that eruptions forced
people to move away due to harmful gases and destructive ash falls,
lahars, and lava ows, but so far we have not detected any clear
evidence of a population inux into adjacent regions during periods of intense volcanic activity. However, after this activity had
ceased, weathering of the deposited ash resulted in the formation
of rich soils and renewed plant growth, likely attracting resettlement of these areas. Based on our data, this appears to have
occurred during the preceramic, at least in some parts of the Middle
Cauca Valley region (Cano et al., 2013). As they weathered, the
fertility of these andisols played an important role in a region
where there is evidence for early plant cultivation (Aceituno and
Lalinde, 2011; Aceituno and Loaiza, 2014a).
Archaeobotanical analysis, particularly starch grain analysis on
stone tools, has documented the use of both local plant resources
and introduced domesticates in the preceramic subsistence economy of the Middle Cauca region. Radiocarbon dates associated with
these stone tools, along with macrobotanical remains, provides a
preliminary history of plant use, cultivar introductions, and the
adoption of horticulture as part of the subsistence strategy. Starch
grains from La Selva and El Jazmn show that local root crops
(Dioscorea sp. and Calathea sp.) and a variety of bean (Phaseolus sp.)
were being used by 8660 55 BP (Aceituno and Loaiza, 2014a),
possibly cultivated in an incipient horticultural system. Tree fruits,
bana (Annona sp.), mora (Rubus sp.), and avocado (P.
such as guana
americana), supplemented the diet, based on macrobotanical re~ ita (Morcote et al.,
mains from El Recreo Cancha and La Montan
2010). Starch grain residues from El Jazmn show that manioc
(Manihot esculenta) was introduced into the region ca. 7590 60
(Aceituno and Loaiza, 2014a). By 7080 50 BP, people were clearly
engaged in cultivation practices, growing maize (Z. mays), along
with manioc, yams, and beans (Aceituno and Lalinde, 2011;
Aceituno and Loaiza, 2014a). Pollen from El Jazmn shows
increased forest clearing around 7000 BP.
These archaeobotanical data suggest that since the early Holocene, human settlers in the Middle Cauca region were collecting,
and possibly cultivating, wild plants in order to optimize the subsistence returns in a humid tropical environment. With this preexisting familiarity with plant resources, people's adoption of
newly introduced domesticates and shift to horticultural activities
likely occurred gradually, without any sudden socio-economic reorientation. Rather than being driven away by frequent volcanic
events, these early horticulturalists took advantage of the soil
fertility to cultivate plants and invest in food production. Research
is ongoing to further investigate the history of plant use in this
region, and understand the changing dynamics of subsistence,
environment, and social organization. A high resolution radiocarbon chronology is essential for establishing the history of these
processes.
5. Conclusions
The western Cordillera Central section of the Middle Cauca region has become a key region in the investigation of preceramic
settlement of Colombia, and the initial transition to food production in the Neotropics. Our recent research in the area has

Please cite this article in press as: Dickau, R., et al., Radiocarbon chronology of terminal Pleistocene to middle Holocene human occupation in the
Middle Cauca Valley, Colombia, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.12.025

R. Dickau et al. / Quaternary International xxx (2014) 1e12

signicantly increased the number of dates and data available for


the region. The chronological sequence obtained from 26 sites indicates a relatively constant and continuous sequence of human
occupation from the terminal Pleistocene to the middle Holocene,
without any signicantly disruption due to volcanic events. Rather
than abandoning the region, people in the Middle Cauca adjusted
culturally to the circumstances of volcanic activity. Of course, this
does not exclude smaller population movements within the region
during volcanic episodes.
The regional patterns of lithic technology, such as the predominance of hoes and grinding stones, and archaeobotanical evidence
of early plant use, including several domesticates, demonstrates the
important role plants played in the adaptive strategies of the preceramic occupants of the Middle Cauca valley. In particular, the
region was likely a major route for the dispersal of crops between
Central America and the rest of South America, the location of
possible domestication of several crop species, and the early
adoption of food production in the American tropics (Piperno and
Pearsall, 1998; Aceituno and Loaiza, 2007, 2014a,b; Dickau, 2008;
Aceituno and Lalinde, 2011; Piperno, 2011). Establishing a solid
chronology in the region forms the basis from which to evaluate
and interpret the timing of these events.
Acknowledgments
Funding was provided by NSF Grant 1049588 to AJR and RD.
Numerous universities, institutions, and associations were instrumental in the successful completion of this project, including the
Dept. of Anthropology at Temple University, the Facultad de Ciengica de Pereira, the
cias Ambientales at the Universidad Tecnolo
Dept. of Anthropology at the Universidad de Antioquia, the
 gicas Universidad de los Andes, the
Departamento de Ciencias Biolo

Dept. of Archaeology at the University of Exeter, the Comite
n Aeropuerto
Departamental de Cafeteros de Caldas, the Asociacio
, the Centro de Museos Universidad de Caldas, Servicio
del Cafe
 gico Colombiano, the Fundacio
 n de Investigaciones
Geolo
gicas Nacionales, and Autopistas del Cafe
. We gratefully
Arqueolo
acknowledge the contribution of the late Dr. Michael Tilst, geologist, to this paper and the Middle Cauca Archaeology Project.
 Botero and palynologist CatThanks also to pedologist Pedro Jose
lez for their insights. The map in Fig. 1 was created by
alina Gonza
squez-Franco for
Juan David Arango, and we thank Susana Vela
helpful discussions regarding Fig. 2. We also would also like to
thank our exceptional excavation crews and the landowners of each
site. The manuscript was improved by comments from two
reviewers.
References
Aceituno, F.J., 2001. Ocupaciones tempranas del bosque tropical subandino en la
cordillera centro-occidental de Colombia. Unpublished Ph.D. dissertation. Facultad de Geografa e Historia, Universidad Complutense de Madrid, Madrid.
Aceituno, F.J., 2002. Interacciones toculturales en el Cauca medio durante el Hol
oceno temprano y medio. Arqueologa del Area
Intermedia 4, 89e113.
Aceituno, F.J., 2010. Nuevos datos para la arqueologa temprana del Cauca Medio: La
mico en la Cuenca del Ro San Eugenio, Cordillera
Pochola, un sitio precera
Central (Colombia). Universidad de Antioquia. Unpublished manuscript.
Aceituno, F.J., Lalinde, V., 2011. Residuos de almidones y el uso de plantas durante el
holoceno medio en el Cauca Medio (Colombia). Caldasia 33, 1e20.
n del bosque en el Cauca medio
Aceituno, F.J., Loaiza, N., 2007. Domesticacio
colombiano entre el Pleistoceno nal y el Holoceno medio. In: British Archaeological Reports International Series 1654. Archaeopress, Oxford.
Aceituno, F.J., Loaiza, N., 2014a. Early and Middle Holocene evidence for plant use
and cultivation in the Middle Cauca River Basin, Cordillera Central (Colombia).
Quaternary Science Reviews 86, 49e62.
Aceituno, F.J., Loaiza, N., 2014b. The role of plants in the early human settlement of
Northwest South America. Quaternary International. http://dx.doi.org/10.1016/
j.quaint.2014.06.027.
Aceituno, F.J., Loaiza, N., Delgado-Burbano, M.E., Barrientos, G., 2013. The initial
human settlement of Northwest South America during the Pleistocene/

11

Holocene transition: synthesis and perspectives. Quaternary International 301,


23e33.
rica: pases andinos Colombia,
Acevedo, E., 1964. Colombia, VVAA. Geografa de Ame
 n, Barcelona.
Ecuador y Venezuela. Montaner y Simo
Berro, J.C., Boom, A., Botero, P.J., Herrera, L.F., Hooghiemstra, H., Romero, F.,
Sarmiento, G., 2001. Multi-disciplinary evidence of the Holocene history of a
cultivated oodplain area in the wetlands of northern Colombia. Vegetation
History and Archaeobotany 10, 161e174.
 n. Boletn Museo del Oro 24, 102e119.
Bray, W., 1989. Cer
amica Buga: Reevaluacio
Bruhns, K.O., 1976. Ancient Pottery of the Middle Cauca Valley, Colombia. Cespedesia 5 (17e18), 101e196.
Bruhns, K.O., 1994. Ancient South America. Cambridge University Press, Cambridge.
gicas en Santuario (Risaralda). Fundacio
n
Cano, M.C., 1995. Investigaciones arqueolo
 gicas Nacionales, Banco de la Repblica, Bogota
.
de Investigaciones Arqueolo
gico del Corredor del Acueducto Red Expresa SurCano, M.C., 1998. Rescate Arqueolo
Oriental, Pereira (Risaralda) Fase II. Empresa de Acueducto y Alcantarillado de
Pereira S.A.-E.S.P., Pereira.
Cano, M.C., 2001. Arqueologa en las Cuencas de los Ros Otn y Consota. FIAN,
Risaralda.
Cano, M.C., 2004. Los Primeros Habitantes de las Cuencas Medias de los Ros Otn y
pez, C.E., Cano, M.C. (Eds.), Cambios Ambientales en Perspectiva
Consota. In: Lo
rica: Ecoregio
n del Eje Cafetero. Universidad Tecnolo
gica de Pereira y
Histo
Programa Ambiental GTZ, Pereira, Colombia, pp. 68e91.
Cano, M.C., 2008. Evidencias Precer
amicas en el Municipio de Pereira: Efectos del
n Temprana de los Bosques Ecuatoriales en el Abanico
Vulcanismo y Colonizacio
nico Pereira-Armenia. In: Lo
 pez, C.E., Ospina, G.A. (Eds.), Ecologa
Fluviovolca
rica: Interacciones Sociedad-Ambiente a Distintas Escalas Socio-TempoHisto
rales. Sociedad Colombiana de Arqueologa, Pereira, pp. 149e168.
pez, C., Me
ndez, R., 2013. Geoarqueologa en ambientes volc
Cano, M.C., Lo
anicos:
impactos ambientales y evidencias culturales en el Cauca Medio (Centro Occidente de Colombia). In: Rubin, J., da Silva, R. (Eds.), Geoarqueologa. PUC Goi
as,
s, Brasil, pp. 227e268.
Goia
 pez, C.E., Realpe, J., 2001. Diez mil an
~ os de huellas culturales en los
Cano, M.C., Lo
suelos del Eje Cafetero, Suelos del Eje Cafetero. Proyecto UTP - GTZ, Pereira,
pp. 184e199.
Cardale de Schrimpff, M., Herrera, L., Bray, W., 2005. The earliest inhabitants. In:
Cardale de Schrimpff, M. (Ed.), Calima and Malagana. Art and Archaeology in
Southwestern Colombia. Pro Calima Foundation, Bogot
a.
~ o), 2001. An
CISAN (Centro De Investigaciones Sociales Antonio Narin
alisis espe, Bogota
.
ciales. Unpublished report prepared for INVIAS. Autopistas del Cafe
nico y el Hombre
Correal, G., 1986. Apuntes sobre el Medio Ambiente Pleistoce
rico en Colombia. In: Bryan, A.L. (Ed.), New Evidence for the Pleistocene
Prehisto
Peopling of the Americas. Center for Study of Early Man. University of Maine,
Orono, Maine, pp. 115e131.
 gicas en los Abrigos
Correal, G., Van der Hammen, T., 1977. Investigaciones Arqueolo
.
Rocosos del Tequendama. Banco Popular, Bogota
mico de
Dickau, R., 2008. El Uso de Maz y Cultgenos de Races en el Precera
 y Colombia: Evidencia de Almidones en Sitios Hmedos Subtropicales
Panama
 pez, C.E., Ospina, G.A. (Eds.), Ecologa Histo
rica:
Premontanos. Pereira. In: Lo
Interacciones Sociedad-Ambiente a Distintas Escalas Socio-Temporales. Sociedad Colombiana de Arqueologa, Pereira, pp. 60e67.
Espinal, L.S., 1990. Zonas de vida de Colombia. Universidad Nacional de Colombia,
Medelln.
gion d'Armenia
Guarn, F., 2008. Etude du Fan Fluvio-Volcanique du Quindio (Re
partement de Ge
ologie et de
Colombie). Unpublished M.Sc. thesis. De
ontologie, Universite
 de Geneve, Switzerland.
Pale
Guarn, F., Gorin, G., Espinosa, A., 2004. A Pleistocene stacked succession of volcaniclastic massows in central Colombia: the Quindio-Risaralda fan. Acta Vulcanologica 16, 109e124.
Guhl, E., 1975. Colombia: Bosquejo de su Geografa Tropical, Tomo I. Instituto
.
Colombiano de Cultura, Bogota
Herd, D.G., 1982. Glacial and Volcanic Geology of the Ruiz - Tolima Complex,
 Publ. Geol. Esp. 8.
Cordillera Central, Colombia. INGEOMINAS, Bogota
Herrera, L., Moreno, M.C., 2011. Informe de un mes de actividades (noviembre 1 a 30
n Aeropuerto del Cafe
. Proyecto de Rescate
de 2011) presentado a la Asociacio
gico Aerocafe
, Palestina, Caldas, Colombia.
Arqueolo

~ a, O., 2009. Metodologa, tipologa cera
mica y el sitio
Herrera, L., Moreno, M.C., Pen
11 El Mirador. Primer volumen del informe nal del Proyecto de Rescate y
gico Aerocafe
 (Palestina, Caldas). Centro de Museos, ViceMonitoreo Arqueolo
 n, Universidad de Caldas, Manizales, Colombia.
rrectora de Proyeccio

~ a, O., 2011. La historia muy antigua del municipio de
Herrera, L., Moreno, M.C., Pen
gico del AeroPalestina (Caldas). Proyecto de Rescate y Monitoreo Arqueolo
. Centro de Museos, Universidad de Caldas, Asociacio
n Aeropuerto del Cafe
, Manizales, Colombia.
puerto del Cafe
, Costa Rica.
Holdridge, L.R., 1967. Life Zone Ecology. Tropical Science Center, San Jose
Hooghiemstra, H., van der Hammen, T., 1993. Late Quaternary vegetation history
and paleoecology of Laguna Pedro Palo (subandean forest belt, Eastern
Cordillera, Colombia). Review of Palaeobotany and Palynology 77, 235e262.
.
IGAC, 1998. Paisajes Fisiogr
acos. ORAM, Bogota
gico, gasoducto de Occidente,
INCIVA, 1995e1996. Proyecto de rescate arqueolo
. Unpublished manuscript.
Mariquita-Yumbo. ECOPETROL, Bogota
INTEGRAL, 1997. Arqueologa de rescate: va alterna de la troncal de Occidente ro
Campoalegre-Estadio Santa Rosa de Cabal. Informe Final. Unpublished report.
 gico de los yacimientos el Jazmn y
Jaramillo, A., Meja, J.C., 2000a. An
alisis palinolo
Guayabito. Departamento de Risaralda, Medelln. Unpublished manuscript.

Please cite this article in press as: Dickau, R., et al., Radiocarbon chronology of terminal Pleistocene to middle Holocene human occupation in the
Middle Cauca Valley, Colombia, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.12.025

12

R. Dickau et al. / Quaternary International xxx (2014) 1e12

gico del yacimiento Campoalegre.


Jaramillo, A., Meja, J.C., 2000b. An
alisis palinolo
dito, Medelln. Unpublished
Departamento de Risaralda. Informe ine
manuscript.
nicas en el territorio Quimbaya: uniJaramillo, L.G., 2008. Sociedades prehispa
sticas, a
reas de actividad y el Complejo Tesorito. Fundacio
 n de
dades dome
gicas Nacionales, Banco de la Repblica, Bogota
.
Investigaciones Arqueolo
Kuhry, P., Salomons, J.B., Riezebos, P.A., Van der Hammen, T., 1983. Paleoecologa de
~ os en el a
rea de la Laguna de Otn-El Bosque. In: Van der
los ltimos 6.000 an
Hammen, T., Perez, P.A., Pinto, P. (Eds.), Studies on Tropical Andean Ecosystems,
La Cordillera Central Colombiana transecto Parque Los Nevados, vol. 1. Cramer,
Vaduz, pp. 227e261.
pez, C., Cano, M.C., 2012. En Torno a los Primeros Poblamientos en el NorLo
rica: Acercamientos desde El Valle Interandino del Magoccidente de Surame
 lica del Per, No. 15.
dalena. Boletn de Arqueologa Ponticia Universidad Cato
pez, C.E., 2008a. Cambios Paisajsticos y Localizacio
n de Evidencias Tempranas en
Lo
pez, C.E., Ospina, G.A. (Eds.), Ecologa
el Valle Medio del Ro Magdalena. In: Lo
rica: Interacciones Sociedad-Ambiente a Distintas Escalas Socio-TempoHisto
gica de Pereira-Sociedad Colombiana de Arqueologarales. Universidad Tecnolo
Universidad del Cauca, Pereira, Colombia, pp. 44e51.
pez, C.E., 2008b. Landscape Development and the Evidence for Early Human
Lo
Occupation in the Inter-Andean Tropical Lowlands of the Magdalena River,
Colombia. Syllaba Press, Miami.
Marchant, R., Behling, H., Berrio, J.C., Cleef, A., Duivenvoorden, J., Hooghiemstra, H.,
Kuhry, P., Melief, B., Schreve-Brinkman, E., Van Geel, B., 2002. Pollen-based
biome reconstructions for Colombia at 3000, 6000, 9000, 12 000, 15 000 and 18
000 14C yr ago: Late Quaternary tropical vegetation dynamics. Journal of
Quaternary Science 17, 113e129.
Marchant, R., Behling, H., Berro, J.C., Cleef, A., Duivenvoorden, J., Hooghiemstra, H.,
Kuhry, P., Melief, R., Schreve-Brinkman, E., van Geel, B., Van der Hammen, T., van
Reenen, G., Wille, M., 2001. Mid- to Late-Holocene pollen-based biome reconstructions for Colombia. Quaternary Science Review 20, 1289e1308.
Melief, A.B., 1985. Late Quaternary Paleoecology of the Parque National los Nevados
Cordillera Central, and Sumapaz Cordillera Oriental areas, Colombia. PhD
dissertation, University of Amsterdam.
Melief, A.B., 1989. Late Quaternary history of vegetational in the Parque Los Nevados
and surrounds (Cordillera Central). In: Van der Hammen, T., Daz-Piedrahita, S.,
Alvarez, V.J. (Eds.), La Cordillera Central Colombiana Transecto Parque de los
Nevados (Segunda Parte). J. Cramer, Berln-Stuttgart, pp. 537e588.
ndez, R., 1997. Atlas de los volcanes activos en Colombia. Ingeominas, Manizales.
Me
ndez, R., Corte
s, G., Cepeda, H., 2002. Evaluacio
n de la amenaza volca
nica
Me
potencial del Cerro Machn (Departamento del Tolima, Colombia). Ingeominas,
Manizales.
n, J., Pen
~ a, P., 2010. Determinacio
n de semillas y restos vegetales
Morcote, G., Beltra
 gicos. Proyecto Arqueolo
 gico Aerocafe
, Bogota
, Colombia. Unpublished
arqueolo
report.
Nieuwenhuis, C.J., 2002. Traces on tropical tools: a functional study of chert artefacts from preceramic sites in Colombia. Archaeological Studies Leiden University. Faculty of Archaeology, Leiden University 9, The Netherlands.
Orozco, J.L., 2001. Las cenizas volc
anicas en el territorio de Pereira y sus alrededores,
VVAA. In: Suelos del eje cafetero. Facultad de Ciencias Ambientales, Coopern Alemana al Desarrollo, Pereira, Colombia, pp. 9e15.
acio
ca. IGAC 6, 8e71.
Oster, R., 1979. Las precipitaciones en Colombia, Colombia Geogra
~ o, D., 1996. Arqueologa de rescate en el gasoducto de occidente Mariquita
Patin
Yumbo PNG-058-95. Volumen I del Informe Final. INCIVA-ECOPETROL, Cali.
~ o, D., Clavijo, A., Go
 mez, A., Pulido, R., Daz, C., 1997. Evidencias paleoindias y
Patin
micas en el Valle del Cauca. Cespedesia 22, 33e95.
cera
Piperno, D.R., 2011. The origins of plant cultivation and domestication in the New
World tropics. Current Anthropology 52, S453eS470.
Piperno, D.R., Pearsall, D.M., 1998. The Origins of Agriculture in the Lowland Neotropics. Academic Press, San Diego.
gico. Sitio Hacienda Ge
nova, Variante Sur.
Restrepo, C.A., 2006. Monitoreo Arqueolo
 n de Investigaciones Arqueolo
gicas
Pereira, Boletn de Arqueologa. Fundacio
Nacionales 21, 3e34.

 gico Fase II. La Romelia e La


Restrepo, C.A., 2008. Informe de Monitoreo Arqueolo
Postrera. Proyecto de Desarrollo Vial Armenia Pereira Manizales, Autopista de
. Instituto Nacional de Vas, Autopistas de Cafe
 S. A., report prepared for
Cafe
 S.A., Bogota
.
INCO-AUTOPISTAS DELCAFE
gico Fase II. Sitio 26 y 67
Restrepo, C.A., 2010. Informe de Monitoreo Arqueolo
Variante Sur de Pereira. Proyecto de Desarrollo Vial Armenia Pereira Manizales,
. Instituto Nacional de Vas, Autopistas de Cafe
 S. A., report
Autopista de Cafe
 S.A., Bogota
.
prepared for INCO-AUTOPISTAS DEL CAFE
gico Condominio campestre Valle de la
Restrepo, C.A., 2012. Componente Arqueolo
Florida. Etapas I y II. Villa Mara, Caldas. Constructora Berln S.A.S. Report.
 gico en el a
rea de inuencia
Restrepo, C.A., 2013a. Informe de Monitoreo Arqueolo
.
del Proyecto de Desarrollo Vial Armenia Pereira Manizales, Autopista de Cafe
 S. A. Documento elaborado para
Instituto Nacional de Vas. Autopistas de Cafe
 S.A.
INCO-AUTOPISTAS DEL CAFE
 gico Finca Miramar, Vereda El Gigante.
Restrepo, C.A., 2013b. Monitoreo Arqueolo
Municipio de Montenegro, Quindo. Informe Final. Ms. Mi Pollo S.A. Final
Report.
gico sitios Los Arrayanes Pk 91150 VillamaraRodrguez, C., 1997. Rescate arqueolo
Caldas y El Pomo Pk 7200 ramal a Manzanares, Fresno-Tolima. Informe Final.
leos (ECOPETROL), Bogota
.
Empresa Colombiana de Petro
nico. Procesos socioculturales antiRodrguez, C., 2002. El valle del Cauca Prehispa
ricas del alto y medio Cauca y la costa Pacca
guos en las regiones neohisto
Colombo-ecuatoriana. Departamento de Historia Facultad de Humanidades
n Taraxacum, Cali, Colombia.
Universidad del Valle, Fundacio
 n de
Rojas, S., Tabares, D., 2000. Consideraciones preliminares para la interpretacio
una historia que empieza a recrearse. Proyecto de desarrollo vial doble calzada
. Fase de Rescate. Excavaciones
Armenia-Pereira-Manizales, Autopista del Cafe
gicas. Informe Final. Instituto Nacional de Vas. CISAN, Bogot
Arqueolo
a. Final
report.
Salomons, J.B., 1989. Paleoecology of volcanic soils in the Colombian Central
Cordillera (Parque Nacional de los Nevados). In: Van der Hammen, T., DazPiedrahita, S., Alvarez, V.J. (Eds.), Studies on Tropical Andean Ecosystems 3: La
Cordillera Central Colombiana transecto Parque de los Nevados. J. Cramer,
Vaduz, pp. 15e216.
Smithsonian Inst., 2013. Global Volcanism Program. Dept. of Mineral Sciences,
National Museum of Natural History (accessed 2014). http://www.volcano.si.
edu/.
Struiver, M., Reimer, P.J., Reimer, R.W., 2005. CALIB 7.0 Radiocarbon Calibration
Program (accessed 2014). http://calib.qub.ac.uk/calib/.
n ro Campoalegre. mundo arcaico en la regio
n
Tabares, D., 2004. Fase I: Prospeccio
del Cauca medio, Colombia. Unpublished report.
 n: arqueologa
Tabares, D., Rojas, S., 2000. Aportes para una historia en construccio
de rescate en la doble calzada Manizales-Pereira-Armenia. INVIAS-CISAN,
Bogota.
nica y Tardiglacial en el
Thouret, J.C., Van der Hammen, T., 1983. La secuencia Holoce
rez, A., Pinto, P. (Eds.), La
Parque los Nevados. In: Van der Hammen, T., Pe
n y
Cordillera Central Colombiana transecto Parque de los Nevados: introduccio
datos iniciales. J. Cramer, Vaduz, pp. 262e276.
Thouret, J., Murcia, A., Salinas, R., Vatin-Perignon, N., 1985. Aspectos volcano
estructurales y dinamismo eruptivo reciente de los volcanes Cerro Bravo y
Nevado del Tolima, Cordillera Central de Colombia. In: Memorias del VI Con, pp. 269e288.
greso Latinoamericano de Geologa, vol. 1. Bogata
, A., Salomons, J.B., 1995. Geologa del
Thouret, J.C., Van der Hammen, T., Juvigne
nico del Ruz-Tolima (Cordillera Central).
cuaternario reciente en el macizo volca
In: Van der Hammen, T., Dos Santos, A. (Eds.), La Cordillera Central Colombiana
transecto Parque de los Nevados (Tercera Parte) Estudios de Ecosistemas Tropandinos. Cramer, Berln-Stuttgart, pp. 183e240.
n Geolo
gica del Paisaje en el Piedemonte del Eje Cafetero
Tilst, M., 2006. La Formacio
pez, C.E., Cano, M.C., Rodrguez, D. (Eds.), Cambios AmbiColombiano. In: Lo
rica: Ecologa Histo
 rica y Cultura Ambiental. Unientales en Perspective Histo
gica de Pereira, Pereira, pp. 79e92.
versidad Tecnolo

Please cite this article in press as: Dickau, R., et al., Radiocarbon chronology of terminal Pleistocene to middle Holocene human occupation in the
Middle Cauca Valley, Colombia, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.12.025

Quaternary International xxx (2014) 1e10

Contents lists available at ScienceDirect

Quaternary International
journal homepage: www.elsevier.com/locate/quaint

14

C data and the early colonization of Northwest South America:


A critical assessment

Miguel Delgado a, b, *, Francisco Javier Aceituno c, Gustavo Barrientos a, b


a

n Antropologa, Facultad de Ciencias Naturales y Museo, Universidad Nacional de La Plata, Paseo del Bosque s/n, B1900FWA La Plata, Argentina
Divisio
Consejo Nacional de Investigaciones Cientcas y T
ecnicas (CONICET), Argentina
c
Departamento de Antropologa, Facultad de Ciencias Sociales y Humanas, Universidad de Antioquia, Calle 67 N 53-108 AA 1226, Medelln, Colombia
b

a r t i c l e i n f o

a b s t r a c t

Article history:
Available online xxx

The aim of this paper is to make a critical appraisal of the available 14C dataset from Northwest South
America (Colombia) corresponding to the Pleistocene/Holocene transition (ca. 12,000e8000 14C years
BP). The rst step in the study was to assemble from both published and unpublished sources an
exhaustive database of 14C dates (n 85), recording data regarding the environmental setting and spatial
coordinates of each site, stratigraphic provenance of the dated samples, material used for dating, and 14C
dating method. After the application of different ltering procedures based on outlier detection techniques, the database was subsequently reduced (n 77). Using uncalibrated and calibrated dates, some
spatial and temporal trends in data distribution were investigated in order to illustrate both the strengths
and weaknesses of the available database. It is concluded that three main features that characterize the
14
C dataset from Northwest South America, namely the very high percentage of 14C measurements made
on charcoal, the almost total disregard of bone as a target sample for dating, and the extremely low
percentage of AMS dates, partially affect both its reliability and comparability. It is suggested that, in
order to use 14C dates as data to make reliable inferences about the timing, pattern, process and tempo of
early exploration and colonization of the study area, work at two different levels would be protably
carried out. In the short term, it would be advisable to develop an extensive and exhaustive program
aimed at redating, with AMS and new sample selection criteria, the more signicant archaeological
assemblages attributable to the Pleistocene/Holocene transition. In the medium to long term, it would be
necessary to implement new research projects specically aimed at obtaining original information about
early human settlement in different geographical areas of the Colombian territory.
2014 Elsevier Ltd and INQUA. All rights reserved.

Keywords:
Radiocarbon data
Paleoenvironmental diversity
Northwest South America
Early human peopling

1. Introduction
Northwest South America, which roughly corresponds to the
current Colombian territory (Fig. 1), is an area from where substantial evidence about Pleistocene human settlement has been
recovered (e.g. Correal and van der Hammen, 1977; Hurt et al.,
1977; Correal, 1986; Ardila, 1991; Niuwenhuis, 1998, 2002;
Gnecco, 2000, 2003; Aceituno, 2001, 2007; Mora and Gnecco,
2003; Gnecco and Aceituno, 2006; Aceituno and Loaiza, 2007;
 pez, 2008; Aceituno et al., 2013). Despite this fact, research
Lo

n Antropologa, Facultad de Ciencias Naturales y


* Corresponding author. Divisio
Museo, Universidad Nacional de La Plata, Paseo del Bosque s/n, B1900FWA La Plata,
Argentina.
E-mail addresses: medelgado@fcnym.unlp.edu.ar, mdelgadoburbano@gmail.
com (M. Delgado).

about the early peopling of this ample and ecologically diverse


territory has never reached the necessary momentum. As a
consequence, there are very few research projects specically
aimed at obtaining information about the timing, pattern, process
and tempo of early exploration and colonization.
In order to lay down the foundations for future research on the
subject, there is an urgent need to critically review a number of
issues regarding evidence about early peopling. Chief among such
issues is the evaluation of the available chronological framework
based on 14C dates obtained at different archaeological contexts
attributable to the Pleistocene/Holocene transition (ca.
12,000e8000 BP). The aim of this paper is to undertake such a
critical review, trying to identify the potential weaknesses of the
existing 14C database, as reviewed by December 2013, corresponding to the targeted time period and suggesting some guidelines for further research. To introduce the problem, we will begin

http://dx.doi.org/10.1016/j.quaint.2014.09.011
1040-6182/ 2014 Elsevier Ltd and INQUA. All rights reserved.

Please cite this article in press as: Delgado, M., et al., 14C data and the early colonization of Northwest South America: A critical assessment,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.09.011

M. Delgado et al. / Quaternary International xxx (2014) 1e10

Fig. 1. Map of Northwest South America showing the main ecogeographic regions
mentioned in the text.

rst with a description of the environmental evolution of Northwest South America during the Pleistocene/Holocene transition,
followed by a brief overview of the archaeological research about
the early peopling of this area.
2. Northwest South America: environmental settings at the
Pleistocene/Holocene transition
From a physiographic point of view, the northwest portion of
the South American subcontinent is a highly diverse area (Fig. 1).
Whereas the Ecuadorian Andes in the south form a continuous
mountain chain, the Colombian Andes split into three branches of
different geological origin, the Western, Central and Eastern
Cordillera, which delimit two wide intermountain river valleys: the
Magdalena and the Cauca (Domnguez, 1988). At both sides of the
Andes lie lowland areas covered with tropical rainforest: the narrow strip along the Pacic coast and the vast Amazon River basin
(Domnguez, 1988). To the north of the latter, extends a large plain
covered with grasses, forested along the rivers. The rest of the area
is made up of rolling savannahs, particularly along the Colombia's
Caribbean coast. Near the equator, the physiographic variants
produce a dramatic climate variation along an altitudinal gradient.
Differences in solar exposure, rainfall, and soils make a vertical
mosaic of markedly different and narrow tiers, except in the
Amazon and eastern plains where the ecosystems are substantially
wider (Gnecco and Aceituno, 2006).
Available palynological, glaciomorphological and diatom evidence allows the reconstruction of the predominant environmental
conditions during the last 18,000 14C years (van der Hammen and
Gonz
alez, 1960; van der Hammen, 1974, 1992; van der Hammen
and Hooghiemstra, 1995; Colinvaux, 1997; Marchant et al., 2001,
lez et al., 2006; Delgado, 2012a). The most con2002, 2004; Ve
spicuous pattern that currently emerges is that Northwest South
America suffered repeated environmental changes of variable intensity over time, inferred from temperature, moisture and rainfall
uctuations as indicated by different climate proxies. Such uctuations may have had the potential to cause signicant ecological
modications resulting from climate-dependent chorological

changes affecting vegetal and animal communities. Recent pollenbased biome reconstructions by Marchant et al. (2001, 2002, 2004)
show that, during the Last Glacial Maximum (LGM) (ca.
18,500e17,500 BP), the lowlands were dominated by grass savannahs, cool mixed forests, and tropical seasonal forests. At mid to
highland settings, a shift from tropical seasonal forests to cool
evergreen forests and cool mixed forests has been registered, while
at locations above 2500 masl there has been a marked increase of
the cool grasslandeshrub biome. In general terms, the vegetation at
this period reects cold and dry conditions.
For the early Lateglacial, between 15,500 and 14,500 BP, the
paleoenvironmental record from mid-lower elevations suggests
that the cool mixed forest biome became more widespread. Cool
mixed forest and cool evergreen forest biomes were recorded at
lower altitudes. The low altitude localities exhibited at that time the
same biomes than today such as grassland savannah, cool mixed
forest and tropical seasonal forest. Overall, the aforementioned
conditions suggest the existence of a cold and dry climate. In
summary, the LGM and the earliest part of the Lateglacial was a
very cold and dry time period, in which some of the investigated
localities had very slow sedimentation rates (Marchant et al., 2002).
Between 12,500 and 11,500 BP, biome reconstructions reveal
environmental conditions relatively similar to those of the previous
period, although with an increased spread of cool evergreen forest
biomes at mid altitudes, thus revealing some climatic amelioration.
This is related to the Guantiva Interstadial (ca. 12,500e11,000 BP),
which is characterized by the increasing of average annual temperature (2 C lower than today) and effective precipitation, as well
as altitudinal movements of the upper forest line (van der Hammen
and Hooghiemstra, 1995). In the Sabana de Bogot
a (Eastern
Cordillera), for instance, there was an expansion of the forest over
ramo (i.e. the ecosystem between the upper forest line and
the pa
the permanent snow line characterized by valleys and plains with a
variety of lakes, peat bogs and wet grasslands intermingled with
shrublands and forest patches; Hofstede et al., 2003), which was
represented by Alnus and vegetation typical of marsh environments
such as lower bushes of the genera Myrica and Symplocos. At
Fquene lake and surrounding areas, the presence of Dodonaea, a
pioneer of bare soil, is a good indicator of this climatic trend (van
der Hammen, 1974, 1992:45), which has correlatives in other
areas of Colombia (Marchant et al., 2002).
The end of the Guantiva Interstadial was marked by the return
of colder and drier conditions associated with the onset of the El
Abra Stadial (ca. 11,000e10,000/9500 BP). Average annual temperatures during the E1 Abra Stadial were 4e6 C lower than today
according to pollen data (van der Hammen and Hooghiemstra,
1995) or 2 e3 C lower than today according to stable isotope
(d13C) data (Boom et al., 2001, 2002; Mora and Pratt, 2002). The
upper forest line during this stadial was some 400e500 m lower
than during the Guantiva Interstadial and some 600e800 m lower
than today (van der Hammen and Hooghiemstra, 1995). Lowered
atmospheric pCO2 and reduced rainfall during this period may have
contributed both to a lower forest line and the colonization of the
tropical Andes by C4 grasses despite prevailing cooler temperatures
(Mora and Pratt, 2002).
, the forest partially disappeared and
In the Sabana de Bogota
ramo, with many open
was replaced by the low bushes of the subpa
ramo grasslands of the family Compositae (van der Hammen and
pa
 nzalez, 1960; van Geel and van der Hammen, 1973; van der
Go
Hammen, 1974, 1978). According to van der Hammen and
Hooghiemstra (1995), the Guantiva-El Abra interval is the
regional equivalent to the European AllerdeYounger Dryas
sequence. Around 10,000 BP the climate ameliorated again, with a
sudden rise in average annual temperature that increased evaporation and caused lakes and swampy areas to dry (van der

Please cite this article in press as: Delgado, M., et al., 14C data and the early colonization of Northwest South America: A critical assessment,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.09.011

M. Delgado et al. / Quaternary International xxx (2014) 1e10

Hammen, 1992). The biome reconstructions for the 9500e8500 BP


interval reveal that there was a notable expansion of mesic biomes
(Marchant et al., 2002). At higher elevations, cool mixed forest
spread at the expense of the cool grasslandeshrub biome. Below
2570 masl, the cool grasslandeshrub biome formed an association
with the cool evergreen forest. At lower altitudes, tropical seasonal
forest and tropical rain forests biomes were present, although some
increasing in the extent of grass savannah and tropical seasonal
forest was detected. According to Marchant et al. (2002, 2004) this
interval was clearly characterized by warmer and wetter conditions
than those of the last part of the Lateglacial.
Changes in the chorology and composition of animal communities during the Pleistocene/Holocene transition are difcult to
establish, as most of the archaeological and paleontological records
come from undated or non-stratied contexts. This particularly
affects the knowledge of relevant aspects regarding Pleistocene
megafauna, whose interactions with humans appear to be poorly
documented (Correal, 1981, 1993; van der Hammen, 1986; Piperno
and Pearsall, 1998; van der Hammen and Correal, 2001; Correal
et al., 2005). Available evidence shows that extinct mammals
including proboscidean gomphotheres (genera Haplomastodon,
Stegomastodon, and Cuvieronius), xenarthrans (genera Gliptodon,
Propraopus, and Eremotherium), and American horses (genus Equus)
coexisted, in certain areas, with still living fauna including cervids
(genera Odocoileus and Mazama), xenarthrans (genus Dasypus),
lagomorphs (genus Sylvilagus), caviomorphs (genera Cavia and
Cuniculus), and cricetids. However, the timing of megafaunal
extinction and the role of humans in the process have yet to be
determined.
3. Archaeological research about early human settlement in
Northwest South America
As in other areas of the Andes, archaeological research in
Northwest South America traditionally focused on the archaeological record of past agricultural communities. As a consequence,
research interest in the problem of the early peopling of the
Colombian territory did not emerge until the 1960s, when ReichelDolmatoff (1965) reported several ndings of lithic artifacts across
the Colombian territory and stressed the importance of investigating a deeper past by indicating the most promising geographic
areas for the search of preceramic cultures. He and the Dutch
geologist T. van der Hammen, who since the 1950s had been
devoted to the investigation of Pleistocene and Holocene paleoenvironmental conditions in the Eastern Cordillera, were pioneers in
the design and implementation of a coherent plan for paleoindian
research (Reichel-Dolmatoff, 1997). The research, focused on the
Sabana de Bogot
a (Eastern Cordillera), was continued by van der
Hammen along with the Colombian archaeologist G. Correal and
the American archaeologist W. Hurt. Between 1967 and the early
1970s, two contexts of Late Pleistocene age were investigated: El
Abra II and Tequendama I. At El Abra II, a rockshelter near the town
 (Cundinamarca), few lithic akes and expedient uniof Zipaquira
facial tools belonging to the so-called Abriense industry or edgetrimmed tool tradition (Hurt, 1977; Correal, 1986) were recovered
at the lower levels (7 and 8) in association with faunal remains of
still living species. Several charcoal samples, mostly small ecks
mixed with dark soil particles, were selected for dating, yielding
Late Pleistocene and Early Holocene ages (Correal et al., 1969; Hurt
et al., 1977; Correal, 1986). At Tequendama (Soacha, Cundinamarca),
Correal and van der Hammen excavated a group of rockshelters. In
the lowers levels of Tequendama I, where two hearths were 14C
dated yielding Late Pleistocene ages, these scholars found faunal
remains similar to those found at El Abra II but associated with
artifacts of a different technology which were labelled as

Tequendamiense industry. At Tequendama, the Tequendamiense


artifacts characterized by the use of allochthonous materials and the
presence of scrapers, thinned akes and a projectile point appear
spatially close to Abriense artifacts (Correal and van der Hammen,
 (Tocancipa
,
1977). Beginning in 1979, G. Correal excavated Tibito
Cundinamarca), an open air site where a number of Abriense stone
tools were recovered associated with bone remains of proboscideans (Haplomastodon sp. and Cuvieronius hyodon), American
horse (Equus sp.) and deer (Odocoileus virginianus). A sample of
bone yielded a Late Pleistocene age (Correal, 1981). At the Sabana
, further investigations in sheltered and open air sites
de Bogota
, Galindo I, Neusa, and Checua lead to the
including Sueva I, Gachala
discovery of contexts characterized by the presence of Abriense
artifacts and faunal remains of still living species, with chronologies
ranging from 12,000 to 8000 BP.
During the 1980s and the early 1990s, several sites were excavated in the Colombian southwest whose ndings were actively
integrated into the discussion regarding early human settlement. At
the Calima region (Western Cordillera), a sub-Andean valley about
1600 masl, two open air sites of Early Holocene agedEl Sauzalito
and El Recreodhave artifact assemblages composed by simple
unifacial akes, hoes, hand stones, hammers, and anvils (Salgado,
an Plateau (Central Cordillera) the San
1988e1990). At the Popay
Isidro site (~1600 masl) has, at a level attributable to the Pleistocene/Holocene boundary, a lithic assemblage composed by thousands of chert artifacts including unretouched and retouched
akes, lanceolate bifaces and handstone tools which is not associated with faunal remains (Gnecco, 1994, 2000, 2003). In addition to
the artifacts, charred seeds of Persea spp. and Erythrina and starch
grains from cf. Xanthosoma, cf. Ipomea, cf. Manihot and cf. Maranta
arundinacea were also identied, as well as other grasses and legumes (Piperno and Pearsall, 1998:200).
Contemporary with the investigations in the Colombian southwest, archaeological research in the Amazon basin led to the dis~ a Roja, an open air site on the middle Caquet
covery of Pen
a River
dated at the beginning of the Early Holocene (Cavelier et al., 1995).
At this site, the lithic assemblage was composed by unifacial akes,
choppers, drills, handstone, milling stone, hammers and anvils
(Cavelier et al., 1995). Thousands of charred seeds and macroremains belonging to different genera of palmae were also recovered at this site (Morcote et al., 1998), along with phytoliths of
Lagenaria spp., Calathea allouia and Cucurbita spp. (Gnecco and
Mora, 1997; Piperno and Pearsall, 1998), indicating the importance of the vegetal resources among early Amazonian people.
During the last decade, some important early sites like Nare, La
~ ones de Bogota
 were
Palestina 1 and 2, San Juan de Bedout, and Pen
explored and subsequently excavated in the lowlands of the Middle
 pez, 2008). At these
Magdalena river valley (for a synthesis, see Lo
stratied open air sites, 14C dated between ca. 10,400 and 8500 BP,
simple akes, plane-convex scrapers and triangular shtail projectile points with straight, oblique or rounded wings, and long thin
tail, made either on chert or quartz were found, with no association
with faunal or botanical remains.
More recently, at the Middle Cauca and the Middle Porce river
valleys (Central Cordillera, ca. 1650e2100 masl), a number of
stratied open air sites whose occupation dates back, in some cases,
to the Late Pleistocene were found (Tabares and Rojas, 2000;
Rodrguez, 2002; Cano, 2004; Castillo and Aceituno, 2006;
Aceituno and Loaiza, 2007; Santos, 2010). At sites such as El Jazmn (Middle Cauca) (Aceituno and Loaiza, 2007) and La Morena
(Middle Porce) (Santos, 2010) the lithic technologydprimarily
consisting of simple akes, axes, hoes, hand stones and milling
basesdis clearly oriented to the exploitation of vegetal resources,
something that is not veried in contemporary sites of the Altiplano Cundiboyacense and the Middle Magdalena River valley.

Please cite this article in press as: Delgado, M., et al., 14C data and the early colonization of Northwest South America: A critical assessment,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.09.011

M. Delgado et al. / Quaternary International xxx (2014) 1e10

Fig. 2. Map of the study region showing the distribution of the archaeological sites
mentioned in the text and in Table 1.

4. The

14

C database

4.1. Data quality assessment and calendar calibration


In order to approach the spatial and temporal distributional
pattern of 14C dates from our study area, we begin by assembling
from both published and unpublished sources an exhaustive
dataset corresponding to the Pleistocene/Holocene transition. For
the purpose of this study, all 14C dates were taken at face value due
to the difculties for an individual-based assessment of each date
imposed by the lack of sufcient information, in most of the original reports, about relevant issues like sample selection criteria,
pre-treatment protocols, and correction for isotopic fractionation.
A total of 85 14C dates allegedly associated with evidence of human

activity at 41 archaeological sites were collected from the existing


literature for further analysis. The archaeological sites consist of
both open air settings (36/41 or 87.8%) and rockshelters (5/41 or
12.2%). The whole sample was examined in search of outliers, either
mild or extreme, on the basis of the calculation of the median, the
interquartile range, and the outlier coefcient (established at 1.5 or
one and a half times the length of the interquartile range) of the
complete distribution of measured 14C ages. This procedure
allowed to detect two mild outliers consisting in a date from El
n (Beta-111972 12,910 60 BP; Salgado, 1998) and a date from
Jorda
El Abra II (GrN-5556 12,400 160 BP; Correal et al., 1969; Hurt
et al., 1977), that were consequently rejected. The rationale for
this decision was to ensure the coherence (i.e. the unity or internal
cohesion) of the sample, under the premise that a continuous or
quasi-continuous distribution of 14C dates is more reliable as a
signal of a colonization phase than a highly punctuated one. A
second screening step consisted of the elimination of all those 14C
dates with exceedingly high sigma values (i.e. laboratory uncertainty), recognized on the basis of an outlier detection analysis
performed as described above. Six extreme values were detected,
ve of them corresponding to El Abra II (Beta-2134 10720 400 BP;
I-6363 9050 470 BP; Beta-2135 8810 430 BP; Beta-2137
8760 350 BP; GrN-82 8670 400 BP; Correal et al., 1969; Hurt
et al., 1977), one to Tequendama I (GrN-6539 10920 250 BP;
~a
Correal and van der Hammen, 1977), and the remaining one to Pen
Roja (GX-17395 9125 250 BP; Cavelier et al., 1995). The resulting
dataset was thus composed of 77 14C dates from the same 41
archaeological considered rst (Table 1; Fig. 2). In this sample, the
one on which all further analyses were carried out, the materials
used for dating were charcoal (70/77 or 92.2%), bone (1/77 or 1.3%),
seeds (1/77 or 1.3%), unspecied organic (1/77 or 1.3%), phytoliths
(1/77 or 1.3%), and undetermined (2/77 or 2.6%). The 14C methods
used for dating were beta-counting (63/77 or 81.8%) and accelerator mass spectrometry (AMS) (14/77 or 18.2%). Decade by decade,
the pace of publication of 14C dates was as follows: 15 in the 1970s
(19.5%), 8 in the 1980s (10.4%), 17 in the 1990s (22.1%), 23 in the
2000s (29.9%), and 14 in the current decade (18.2%). No statistically
signicant differences between the number of dates published in
consecutive decades were observed (test of difference between
percentages, two sided distribution, a 0.05).

Table 1
14
C dates from Northwest South America at the Pleistocene/Holocene transition.
N

Site

Altitude
(masl)

Sample

Code

14

1
2
3


Tibito
El Abra II
Tequendama I

2590
2570
2566

bone
organic
charcoal

GrN-9375
GrN-5941
GrN-6270

11740
11210
10730

110
90
105

BeCa
BeC
BeC

open air
Altiplano CB
rockshelter Altiplano CB
rockshelter Altiplano CB

Tequendama I

2566

charcoal

GrN-6505

10590

90

BeC

rockshelter Altiplano CB

Tequendama I

2566

charcoal

GrN-6731

10460

130

BeC

rockshelter Altiplano CB

6
7
8
9
10
11
12
13
14

Nare
175
Nare
175
La Palestina 2
130
Nare
175
San Juan de Bedout
130
La Palestina 2
130
PIII0I-52
1524
La Palestina 2
130
Tequendama I
2566

charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal

Beta-146798
Beta-70040
Beta-40855
Beta-70040
Beta-40852
Beta-123566
Beta 205293
Beta-123565
GrN-6210

10400
10400
10400
10350
10350
10300
10260
10260
10250

60
40
90
60
90
70
50
70
95

AMSb
AMS
BeC
BeC
BeC
BeC
AMS
BeC
BeC

open air
open air
open air
open air
open air
open air
open air
open air
rockshelter

C date Sigma Technique Setting

Regionc

Middle Magdalena
Middle Magdalena
Middle Magdalena
Middle Magdalena
Middle Magdalena
Middle Magdalena
Central Cordillera
Middle Magdalena
Altiplano CB

15 La Palestina 2
16 Tequendama I

130
2566

charcoal
charcoal

Beta-40854
GrN-7114

10230
10150

80
150

BeC
BeC

open air
Middle Magdalena
rockshelter Altiplano CB

17 Tequendama I

2566

charcoal

GrN-7113

10140

100

BeC

rockshelter Altiplano CB

References
Correal (1981)
Hurt et al. (1977)
Correal and van der
Hammen (1977)
Correal and van der
Hammen (1977)
Correal and van der
Hammen (1977)
pez (2008)
Lo
pez (2008)
Lo
pez (2008)
Lo
pez (2008)
Lo
pez (1989)
Lo
pez (2008)
Lo
Otero and Santos (2006)
pez (2008)
Lo
Correal and van der
Hammen (1977)
pez (2008)
Lo
Correal and van der
Hammen (1977)
Correal and van der
Hammen (1977)

Please cite this article in press as: Delgado, M., et al., 14C data and the early colonization of Northwest South America: A critical assessment,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.09.011

M. Delgado et al. / Quaternary International xxx (2014) 1e10


Table 1 (continued )
Altitude
(masl)

Sample

Code

14

18 El Guatn
19 Tequendama I

1440
2566

charcoal
charcoal

Beta-325213
GrN-6732

10130
10130

50
150

AMS
BeC

open air
Central Cordillera
rockshelter Altiplano CB

20
21
22
23
24
25

El Jazmn
Sueva I
La Morena
San Isidro
San Isidro
Tequendama I

1650
2690
1650
1690
1690
2566

charcoal
charcoal
charcoal
charcoal
seed
charcoal

Ua-24497
GrN-8111
Beta-245566
Beta-65878
Beta-93275
GrN-6730

10120
10090
10060
10050
10030
9990

70
90
60
100
60
100

BeC
BeC
BeC
BeC
AMS
BeC

open air
rockshelter
open air
open air
open air
rockshelter

Central Cordillera
Altiplano CB
Central Cordillera
Central Cordillera
Central Cordillera
Altiplano CB

26 La Palestina 1

130

charcoal

n.d.

9820

115

BeC

open air

Middle Magdalena

n
27 El Jorda
28 Tequendama I

2640
2566

charcoal
charcoal

Beta-116764
GrN-7115

9760
9740

160
135

BeC
BeC

open air
Central Cordillera
rockshelter Altiplano CB

29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46

1270
1650
2000
1500
1500
1500
1500
1000
1690
1600
1758
2570
1000
2570
1700
1500
1500
141

charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
n.d.
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal

Beta-121972
Beta-245564
Beta-146613
Beta-23476
Beta-23476
Beta-23475
Beta-23475
CEDAD LTL 4267A
Beta-65877
Beta-87188
GrN-8448
GrN-5661
CEDAD LTL 4845A
GrN-5746
LTL-4223A
Beta-8.441
Beta-18441
Beta-52964

9730
9680
9680
9670
9670
9600
9600
9542
9530
9490
9360
9340
9333
9325
9312
9300
9300
9250

100
60
100
150
100
110
100
50
100
110
45
40
65
100
55
100
100
140

BeC
BeC
BeC
BeC
BeC
BeC
BeC
AMS
BeC
BeC
BeC
BeC
BeC
BeC
BeC
BeC
BeC
BeC

open air
open air
open air
open air
open air
open air
open air
open air
open air
open air
rockshelter
rockshelter
open air
rockshelter
open air
open air
open air
open air

Central Cordillera
Central Cordillera
Central Cordillera
Western Cordillera
Western Cordillera
Western Cordillera
Western Cordillera
Central Cordillera
Central Cordillera
Central Cordillera
Altiplano CB
Altiplano CB
Central Cordillera
Altiplano CB
Central Cordillera
Central Cordillera
Western Cordillera
Amazon Basin

nova
47 Ge
~ ita
48 La Montan
~ a Roja
49 Pen

1000
1000
141

charcoal
charcoal
charcoal

Beta-355217
Beta-355214
Beta-64602

9230
9230
9160

40
50
90

AMS
AMS
BeC

open air
open air
open air

Central Cordillera
Central Cordillera
Amazon Basin

50
51
52
53
54
55
56
57
58
59
60
61
62
63

1000
1700
2570
1650
1000
110
1291
2689
1600
1600
1611
1600
1600
1341

charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal
charcoal

Beta-72375
LTL5436A
Beta-5710
Beta-95061
Beta-114687
Beta-26018
Beta-306257
GrN-16346
X-23803
X-23805
Ua-24498
Ua-24499
X-23804
Beta-285871

9120
9047
9025
9020
8990
8750
8740
8740
8712
8704
8680
8680
8674
8550

90
45
90
60
80
160
50
60
60
56
55
60
61
60

BeC
BeC
BeC
BeC
BeC
BeC
AMS
BeC
BeC
BeC
BeC
BeC
BeC
BeC

open air
open air
rockshelter
open air
open air
open air
open air
open air
open air
open air
open air
open air
open air
open air

Central Cordillera
Central Cordillera
Altiplano CB
Central Cordillera
Central Cordillera
Western Cordillera
Central Cordillera
Altiplano CB
Central Cordillera
Central Cordillera
Central Cordillera
Central Cordillera
Central Cordillera
Central Cordillera

Restrepo (2013)
Correal and van der
Hammen (1977)
Aceituno and Loaiza (2007)
Correal (1979)
Santos (2010)
Gnecco (2000)
Gnecco (2000)
Correal and van der
Hammen (1977)
CAIN-OCENSA (1997)
pez (2008)
in Lo
Salgado (1998)
Correal and van der
Hammen (1977)
Cano (2004)
Santos (2010)
Tabares and Rojas (2000)
Bray et al. (1988)
Bray et al. (1988)
Bray et al. (1988)
Bray et al. (1988)
Restrepo (2013)
Gnecco (2000)
Rodrguez (2002)
Correal (1979)
Hurt et al. (1977)
Restrepo (2013)
Hurt et al. (1977)
Aceituno ms.
Bray et al. (1988)
Bray et al. (1988)
Cavelier et al. (1995);
Gnecco (2000)
Restrepo (2013)
Restrepo (2013)
Cavelier et al. (1995);
Gnecco (2000)
Castillo and Aceituno (2006)
Aceituno ms.
Hurt et al. (1977)
INTEGRAL (1997)
Castillo and Aceituno (2006)
Herrera et al. (1992)
Restrepo (2013)
Pinto (2003)
Aceituno and Loaiza (2014)
Aceituno and Loaiza (2014)
Aceituno and Loaiza (2007)
Aceituno and Loaiza (2007)
Aceituno and Loaiza (2014)
Herrera et al. (2011)

141
1341

n.d.
charcoal

Beta-64601
Beta-290954

8510
8480

110
40

BeC
AMS

open air
open air

Amazon Basin
Central Cordillera

Llanos (1997)
Herrera et al. (2011)

116

charcoal

Beta-181064

8480

40

AMS

open air

Middle Magdalena

pez (2008)
Lo

2391
1950
3350
920
3340

charcoal
charcoal
charcoal
charcoal
charcoal?

Beta-146609
Beta-93154
Beta-21060
Beta 231479
GrN 12068

8430
8380
8370
8340
8320

100
90
90
40
80

BeC
BeC
BeC
AMS
BeC

open air
open air
rockshelter
open air
open air

Central Cordillera
Central Cordillera
Altiplano CB
Central Cordillera
Altiplano CB

Tabares and Rojas (2000)


INTEGRAL (1997)
Rivera (1991)
Cardona et al. (2007)
Herrera et al., 1992

1000
2600
1600
1700
141
1341

charcoal
charcoal
charcoal
charcoal
phytoliths
charcoal

Beta-325216
A-03
CSIC-1987
Ua-24499
UCR-3419; CAMS-27728
Beta-283582

8240
8200
8136
8095
8090
8030

40
110
65
55
60
80

AMS
BeC
BeC
BeC
AMS
BeC

open
open
open
open
open
open

Central Cordillera
Altiplano CB
Central Cordillera
Central Cordillera
Amazon Basin
Central Cordillera

Restrepo (2013)
Groot (1992)
Aceituno and Loaiza (2007)
Aceituno and Loaiza (2007)
Piperno and Pearsall (1998)
Herrera et al. (2011)

N

64
65
66
67
68
69
70
71
72
73
74
75
76
77
a
b
c

Site

66PER001
La Morena
Salento 24
Sauzalito
Sauzalito
Sauzalito
Sauzalito
La Trinidad Corte I
San Isidro
La Selva Risaralda
Gachal
a
El Abra II
La Trinidad Corte II
El Abra II
La Pochola
Sauzalito
Sauzalito
~ a Roja
Pen

Sitio 045
La Pochola
El Abra II
El Jazmn
Sitio 021
El Recreo
Nuevo Sol
Galindo I
La Selva
La Selva
La Pochola
La Selva
La Selva
39 El Recreo
Cancha
~ a Roja
Pen
39 El Recreo
Cancha
~ ones de
Pen

Bogota
Salento 21
El Antojo
Neusa
PIII0P-59
mo de
Para
~ a Negra
Pen
La Chillona
Checua
San Germ
an II
La Pochola
~ a Roja
Pen
39 El Recreo
Cancha

C date Sigma Technique Setting

air
air
air
air
air
air

Regionc

References

Beta Counting.
Accelerator Mass Spectrometry.
Altiplano Cundiboyacense.

Please cite this article in press as: Delgado, M., et al., 14C data and the early colonization of Northwest South America: A critical assessment,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.09.011

M. Delgado et al. / Quaternary International xxx (2014) 1e10

Due to the lack of sufcient contextual information, no attempts


to associate 14C dates from any particular site through the calculation of pooled means were made prior to calibration. The calendar
calibration of the putative conventional 14C dates was performed
using CALPAL_A (Cologne Radiocarbon Calibration & Palaeoclimate
Research Package), a software package that allows the simultaneous processing of a large bulk of data and the comparison of the
summed probability distribution with paleoclimatic curves from
different proxies. The calibration curve chosen in this study was the
ris, 2008), which produces reCalPal-2007-Hulu (Weninger and Jo
sults that are identicaldfor all practical purposesdto the recently
published and internationally recommended INTCAL09-calibration
ller et al., 2012).
(Bradtmo
4.2. Spatial and temporal trends in

14

C data

The geographic patterning of early dates was assessed by means


of a raster surface map using the kriging algorithm. We used XYZ
coordinate data where X and Y are, respectively, the longitude and
latitude of each archaeological site and Z is the value corresponding
to the earliest 14C date registered at each site. As Fig. 3 shows, the
earliest sites are distributed in the Altiplano Cundiboyacense
(Eastern Cordillera) and the Middle Magdalena River basin. To the
west, the occupation of the Middle and Western Cordillera and of
the Cauca River valley seems to have been a later phenomenon. To
the east, the only early site discovered so far in the Amazonas Basin,
~ a Roja, has a record of occupation that begins at the Early
Pen
Holocene.
Radiocarbon dates corresponding to the Pleistocene/Holocene
transition come from sites distributed at three main altitudinal
oors: 0e500 masl, 1000e2000 masl, and 2500e3000 masl (Fig. 4).
It is remarkable that the two oldest dates, both exceeding 11,000 BP,
were registered at sites corresponding to the third altitudinal oor,
 and El Abra, Altiplano Cundiboyacense).
at about 2600 masl (Tibito
The occupation of the lowest oor, mainly represented at the
Middle Magdalena River basin, seems to begin later, at around
10,500 BP, at Nare, La Palestina 2, and San Juan de Bedout. The

Fig. 3. Raster surface map (kriging algorithm), superimposed to the map of the study
region, showing the spatio-temporal trends in the distribution of the earliest 14C date
registered at each archaeological site.

Fig. 4. Scatterplot with histograms showing the distribution of uncalibrated 14C dates
from Northwest South America against the altitude of the respective archaeological
sites.

second oor, located between 1000 and 2000 masl in the Central
and Western Cordillera, has a record of occupation that seems to
begin slightly later, around 10,100 BP.
Regarding the temporal pattern of distribution of the 14C dates
corresponding to the Pleistocene/Holocene transition, both the
uncalibrated and the calibrated sets (2 sigma) exhibit the same
features (Figs. 5 and 6), i.e. a weak and punctuated signal before
11,000 BP and a stronger and continuous signal after that date,
 and
particularly in Late Pleistocene times. The only date from Tibito
the earliest date from El Abra II (11,740 110 BP and 11,210 90 BP,
respectively) are disconnected from the rest of the dates,
notwithstanding the fact that they do not behave as outliers in a
statistical sense. This makes these samples obvious candidates for
redating in order to conrm or reject the estimated 14C ages.
In terms of the environmental scenario in which the early
peopling of Northwest South America occurred, if the two pre11,000 BP dates are accepted, then the colonization of the study
area likely began at the Guantiva Interstadial, a period in which the
Altiplano Cundiboyacense had relatively warm and humid

Fig. 5. Range plot displaying error bars (2 sigmas) corresponding to


Pleistocene/Holocene transition from Northwest South America.

14

C dates of the

Please cite this article in press as: Delgado, M., et al., 14C data and the early colonization of Northwest South America: A critical assessment,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.09.011

M. Delgado et al. / Quaternary International xxx (2014) 1e10

Fig. 6. (Top) summed probability distribution plot of calibrated ages (2 sigmas) from
Northwest South America; (bottom) paleoclimate proxies: GISP2 d18O (Meese et al.,
1997) and Cariaco Varve Greyscale (Hughen et al., 1998). YDC Younger Dryas
Chronozone.

conditions that allowed the spread of a cool grassland/shrub biome


and an ascendant movement of the upper forest line (Fig. 7). If both
dates are deemed provisionally dubious, then the colonization
process began during the El Abra Stadial, contemporary with the
Younger Dryas Chronozone (12,900e11,600 cal BP; Newby et al.,
2005) (Fig. 7). In either case, it was during this cold pulse when
the colonizing population rst reached an indisputably clear
archaeological visibility represented by a diversity of contexts
deposited at different environmental settings in the Altiplano
Cundiboyacense and the Middle Magdalena River valley.
5. Discussion
In the last decade, there is a growing recognition that 14C dates,
beside their usefulness for establishing the temporal position of
past specic events, can be used as data to discuss a number of
archaeological problems (Rick, 1987). In particular, datasets

Fig. 7. Histogram of uncalibrated

collected from large geographic areas are important sources of information about population processes such as demographic change
and spatial dispersal (see, among others, Housley et al., 1997;
Anderson and Faught, 2000; Bocquet-Appel and Demars, 2000;
 ska and Pazdur,
Gkiasta et al., 2003; Gamble et al., 2004; Michczyn
2004; Barrientos and Perez, 2005; Chamberlain, 2006; Hamilton
and Buchanan, 2007; Buchanan et al., 2008, 2011; Riede, 2009;
Collard et al., 2010; Peros et al., 2010; Steele, 2010; Wilmshurst
ller et al., 2012;
et al., 2011; Bamforth and Grund, 2012; Bradtmo
Williams, 2012; cf. Surrovell and Brantingham, 2007; Delgado,
2012b; Barrientos and Masse, 2014). In this line of research a major concern is, or should be, the quality of the databases in terms of
some critical archaeological (e.g. associational, stratigraphic) and
chronometric (e.g. pre-treatment and measurement) criteria
(Waterbolk, 1971; Pettitt et al., 2003; Vermeersch, 2005).
In the case of Northwest South America, there are some practical
difculties in thoroughly assessing, on a case by case basis, the
quality of the 14C dates due to the generalized incompleteness of
the bibliographic sources regarding relevant issues including
sample selection criteria, degree of association between dated
samples and the archaeological phenomena they are intended to
represent (i.e. sample-event association), pre-treatment protocols,
correction for isotopic fractionation (in the case of the only bone
sample dated so far), and 14C measurement techniques and protocols. Notwithstanding this fact, some considerations can be made
regarding the dataset as a whole.
Firstly, it is quite remarkable the overwhelming majority of 14C
dates obtained on charcoal samples and the almost null representation of dates obtained on bone. Both kinds of materials, used as
samples for dating, have advantages and problems. On the one side,
charcoal is often preferred over bone due to (i) the higher vulnerability of the latter to contamination with substances such as humic
and fulvic acids, polyphenols, polysaccharides, lignins, and
degraded collagen, (ii) the decay of collagen with time, and (iii) the
much least complex pre-treatment required by charcoal samples
ris et al.,
compared with bone collagen (Stafford et al., 1987; Jo
2003). On the other side, the main advantages of bone collagen
over charcoal reside in the facts that (i) the former is a short-lived
material whereas wood charcoal can be affected by the so-called

14

C dates against a paleoenvironmental chart (by region) of Northwest South America at the Pleistocene/Holocene transition.

Please cite this article in press as: Delgado, M., et al., 14C data and the early colonization of Northwest South America: A critical assessment,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.09.011

M. Delgado et al. / Quaternary International xxx (2014) 1e10

old wood effect (Taylor, 1987; Bowman, 1990) and (ii) bone (both
faunal and human) usually has, compared to charcoal, a higher
reliability as a marker of human activity, particularly when an
adequate knowledge of taphonomic factors and processes operrrez, 2004; Borrero, 2008,
ating at each locality is available (Gutie
2009). Moreover, most of the problems regarding bone contamination have been overcome due to renements in pre-treatment
techniques (Hedges and Van Klinken, 1992; Yuan et al., 2000;
Bronk Ramsey et al., 2004) and the increasing use of AMS dating
over the last 30 years (Gillespie et al., 1984; Gowlett, 1987). The
latter is a method based on the counting of 14C atoms instead of 14C
decay (beta particles) that requires very small samples (i.e., less
than a milligram, approximately a thousand times less material
than is required from beta-counting methods; Currie, 2004;
Gowlett, 1987) and allows, when necessary, for the dating at the
molecular level (i.e. individual bone collagen amino acids) (Stafford
et al., 1987; Stafford, 1990), thus enhancing the suitability and
reliability of bone as sample for 14C dating. In general terms, then, a
dataset composed by 14C dates obtained mostly on bone samples
with good taphonomic control can be considered (all other things
being equal) as having a stronger sample-event relationship than a
one like ours, mostly composed by dates obtained on charcoal.
Secondly, less than 20% of the dates were obtained with AMS,
which is a very low gure considering that more than 70% of the
dates were published well after the inception and popularization of
this method in the late 1980s. This suggests that, in order to full
the weight requirements of traditional beta-counting methods
(5e20 g), most of the wood charcoal samples used for dating were
likely a composite of ecks, which increases the uncertainty about
the effect of the old wood problem and the overall strength of the
sample-event relationship.
Thirdly, almost the 40% of the 14C dates that yielded Pleistocene
ages were obtained in the 1970s. While alone is not enough to cast
doubt on the reliability of such dates, it raises questions about the
comparability between them and those more recently obtained
dates owing to the advances experienced by 14C dating over the last
decades, particularly regarding pre-treatment protocols. It is signicant that the ve oldest dates in our database, with 14C ages
ranging between 11,740 110 and 10,460 130 BP, belong to this
group (Table 1). They come from the three classic early sites of the
 (Altiplano Cundiboyacense): Tibito
 , El Abra II, and
Sabana de Bogota
Tequendama I. While the lower levels from El Abra II and Tequendama I have been repeatedly dated, both cases exhibiting a
 has only a
considerable dispersion of the estimated 14C ages, Tibito
single date, the only one in our database obtained from a bone
sample (Correal, 1981). As it is widely recognized, the age of any
single event cannot be considered as reliably established on the
basis of a single 14C date, it being necessary to have a number of
dates in statistical agreement (Geyh and De Maret, 1982). In the
, there is an even more pressing need to redate the
case of Tibito
archaeological level, as the only available date places it as the oldest
site in the region yet discovered.
On the basis of the above considerations, we believe that the
spatial and temporal trends in 14C data discussed in this paper have
to be taken with caution until there is more and better information.
In the near future, however, such an improvement is unlikely to
occur due to the near absence of research projects specically
aimed at obtaining novel information about the early peopling of
Northwest South America. The current lack of a critical mass of local
archaeologists interested in this problem is a major impediment for
the rapid growth of such an important area of scientic research. In
this context, the development of an extensive and exhaustive
program aimed at redating, with AMS and new sample selection
criteria, the more signicant archaeological assemblages attributable to the Pleistocene/Holocene transition appears as a much

desirable and accomplishable goal in the short term. This may also
contribute to augment the degree of comparability between the
Colombian database and those from other regions of South America, particularly the Southern Cone, where AMS dating, taphonomic
control, bone as preferred sample, and calendar calibration of large
datasets are becoming established as standard procedures (e.g.
rez, 2003; Gutie
rrez, 2004; Barrientos and Perez, 2005;
Rubinos Pe
Borrero, 2008, 2009; Steele and Politis, 2009; Barrientos and Masse,
2014). Interregional comparisons at the continental level are relevant in order to detect differences in the date of initial colonization
and subsequent occupation patterns as well as to estimate dispersal
rates, which are important tools for the understanding of the demographic and socio-ecological properties of early South American
populations.
6. Concluding remarks
Archaeological work carried out during the last ve decades in
Northwest South America allowed the recovery of substantial evidence about human settlement during the Pleistocene/Holocene
transition, key to understanding the early human entry into the
area and the subsequent dispersal through the subcontinent.
However, it is clear that new archaeological evidence and more
reliable radiometric dating are necessary in order to expand and
rene our current knowledge about such important issues as biocultural diversity, economic strategies, and rate and pathways of
population dispersal of the rst settlers. In this context, work at two
different levels would be protably carried out: (i) in the short
term, it would be advisable to develop an extensive and exhaustive
program aimed at redatingdwith AMS and new sample selection
criteriadthe more signicant archaeological assemblages attributable to the Pleistocene/Holocene transition, particularly those
that yielded very old dates like El Jord
an and El Abra II; (ii) in the
medium to long term, it would be necessary to implement new
interdisciplinary research projects specically aimed at obtaining
original information from different geographical areas. Undoubtedly, the whole enterprise would require higher levels of funding,
scholarly cooperation, and academic involvement than that which
characterized research about the early peopling of the Colombian
territory.
Acknowledgements
 bal Gnecco who provided literature
We are indebted to Cristo
and opportune advice. Part of the research conducted for this paper
n de Investigaciones Arqueolo
gicas
was supported by the Fundacio
Nacionales (FIAN), Colombia, through a research grant (grant
number 4092010) awarded to the senior author. This paper is
dedicated to the memory of Luis F. Delgado (1937e2013) loving and
devoted father.
References
Aceituno, F.J., 2001. Ocupaciones tempranas del bosque tropical subandino en la
Cordillera Centro-Occidental de Colombia (Unpublished PhD dissertation).
Facultad de Geografa e Historia, Universidad Complutense de Madrid, Madrid.
n Andina del
Aceituno, F.J., 2007. Poblamiento y variaciones culturales en la regio
rica en la transicio
n Pleistoceno/Holoceno. In:
noroccidente de Surame
 n, C., Pupio, A., Gonza
lez, I., Flegenheimer, N., Fre
re, M. (Eds.), Arqueologa
Bayo
de las Pampas. Sociedad Argentina de Antropologa, Buenos Aires (Buenos
Aires), pp. 15e38.
n del bosque en el Cauca Medio
Aceituno, F.J., Loaiza, N., 2007. Domesticacio
Colombiano entre el Pleistoceno nal y el Holoceno medio. In: British
Archaeological Reports, International Series, vol. 1654. Archaeopress, Oxford.
Aceituno, F.J., Loaiza, N., 2014. The role of plants in the early human settlement of
Northwest South America. Quaternary International. http://dx.doi.org/10.1016/
j.quaint.2014.06.027.

Please cite this article in press as: Delgado, M., et al., 14C data and the early colonization of Northwest South America: A critical assessment,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.09.011

M. Delgado et al. / Quaternary International xxx (2014) 1e10


Aceituno, F.J., Loaiza, N., Delgado Burbano, M.E., Barrientos, G., 2013. The initial
settlement of Northwest South America during the Pleistocene/Holocene
transition: synthesis and perspectives. Quaternary International 301, 23e33.
Anderson, D.G., Faught, M.K., 2000. Paleoindian artifact distributions: evidence and
implications. American Antiquity 74, 507e513.
Ardila, G., 1991. The peopling of Northern South America. In: Bonnichsen, R.,
Turnmire, K.L. (Eds.), Clovis Origins and Adaptations. Center for the Study of the
First Americans. Oregon State University, Corvallis, pp. 261e282.
Banforth, D.B., Grund, B., 2012. Calibration curves, summed probability distributions, and early paleoindian population trends in North America. Journal of
Archaeological Science 39, 1768e1774.
Barrientos, G., Perez, S.I., 2005. Was there a population replacement during the late
mid-Holocene in the southeastern Pampas of Argentina? Archaeological evidence and paleoecological basis. Quaternary International 132, 95e105.
Barrientos, G., Masse, B., 2014. The archaeology of cosmic impact: lessons from two
Mid-Holocene Argentine case studies. Journal of Archaeological Method and
Theory 21, 134e211.
Bocquet-Appel, J.P., Demars, P.Y., 2000. Neanderthal contraction and modern human colonization of Europe. Antiquity 74, 544e552.
Boom, A., Mora, G., Cleef, A.M., Hooghiemstra, H., 2001. High altitude C4 grasslands
in the northern Andes: relicts from glacial conditions? Review of Paleobotany
and Palynology 115, 147e160.
, J.S., 2002. Co2- and
Boom, A., Marchant, R., Hooghiemstra, H., Sinninghe Damste
temperature controlled altitudinal shifts of C3- and C4-dominated grasslands
allows reconstructions of palaeoatmospheric pCO2. Paleogeography, Paleoclimatology, Paleoecology 177, 151e168.
Borrero, L.A., 2008. Extinction of Pleistocene megamammals in South America: the
lost evidence. Quaternary International 185, 69e74.
Borrero, L.A., 2009. The elusive evidence: the archaeological record of the South
American extinct megafauna. In: Haynes, G. (Ed.), American Megafaunal Extinctions at the End of the Pleistocene. Springer Science, Dordrecht,
pp. 145e168.
Bowman, S., 1990. Radiocarbon Dating. University of California Press, Berkeley.
ler, M., Pastoors, A., Weninger, B., Weniger, G., 2012. The repeated
Bradtmo
replacement model e rapid climatic change and population dynamics in late
Pleistocene Europe. Quaternary International 247, 38e49.
Bray, W., Herrera, L., Schrimpff, M., 1988. Report on the 1984 Field Season. ProCaologisch-ethnologisches Projekt im Westlichen Kolumbien/Sdalima. Archa
merika No 5:2e42. Periodische Publikation der Vereinigung Pro Calima, Basel.
Bronk Ramsey, C., Higham, T., Bowles, A., Hedges, R., 2004. Improvements to the
pretreatment of bone at Oxford. Radiocarbon 146, 155e163.
Buchanan, B., Collard, M., Edinborough, K., 2008. Paleoindian demography and the
extraterrestrial impact hypothesis. Proceedings of the National Academy of
Sciences of the United States of America 105, 11651e11654.
Buchanan, B., Hamilton, M., Edinborough, K., O'Brien, M.J., Collard, M., 2011.
A comment on Steele's (2010) radiocarbon dates as data: quantitative strategies
for estimating colonization front speeds and event densities. Journal of
Archaeological Science 38, 2116e2122.
Cano, M., 2004. Los primeros habitantes de las cuencas medias de los ros Otn y
pez, C., Cano, M. (Eds.), Cambios Ambientales en Perspectiva
Consota. In: Lo
rica. Ecorregio
n del Eje Cafetero, vol. I. Universidad Tecnolo
gica de Pereira,
Histo
Programa Ambiental GTZ, Pereira, pp. 68e91.
Cardona, Luis C., Luis, E. Nieto, Jorge, I. Pino, 2007. Del Arcaico a la Colonia. Conn del paisaje y cambio social en el Porce Medio. Informe nal. Unistruccio
versidad de Antioquia-Empresas de Medelln, Medelln. Unpublished.
Castillo, N., Aceituno, F., 2006. El bosque domesticado, el bosque cultivado: un
proceso milenario en el valle medio del rio Porce en el noroccidente Colombiano. Latin American Antiquity 17, 561e578.
Cavelier, I., Rodrguez, C., Herrera, L., Morcote, G., Mora, S., 1995. No solo de la caza
n del bosque amazo
nico, Holoceno temprano. In:
vive el hombre: Ocupacio

rica
Cavelier, I., Mora, S. (Eds.), Ambito
y Ocupaciones Tempranas de la Ame
 n Erigaie, Instituto Colombiano de Antropologa, Bogota
,
Tropical. Fundacio
pp. 27e44.
Chamberlain, A.T., 2006. Demography in Archaeology. Cambridge University Press,
Cambridge.
Colinvaux, P., 1997. Pleistocene landscapes: the ice-age Amazon and the problem of
diversity. The Review of Archaeology 19, 1e10.
Collard, M., Buchanan, B., Hamilton, M.J., O'Brien, M.J., 2010. Spatiotemporal dynamics of the Clovis-Folsom transition. Journal of Archaeological Science 37,
2513e2519.
gicas en los Abrigos Rocosos de Nemoco
n
Correal, G., 1979. Investigaciones Arqueolo
n de Investigaciones Arqueolo
gicas Nacionales, Bogot
y Sueva. Fundacio
a.
nica en Colombia.
Correal, G., 1981. Evidencias Culturales y Megafauna Pleistoce
n de Investigaciones Arqueolo
gicas Nacionales, Bogota
.
Fundacio
nico y el hombre
Correal, G., 1986. Apuntes sobre el medio ambiente pleistoce
rico en Colombia. In: Bryan, A.L. (Ed.), New Evidence for the Pleistocene
prehisto
Peopling of the Americas. Center for the Study of the Early Man, University of
Main, Orono, pp. 115e131.
nicas y megafauna en
Correal, G., 1993. Nuevas evidencias culturales pleistoce
Colombia. Boletn de Arqueologa 8, 3e12.
gicas en los abrigos
Correal, G., van der Hammen, T., 1977. Investigaciones arqueolo
.
rocosos del Tequendama. Biblioteca Banco Popular, Bogota
Correal, G., van der Hammen, T., Lerman, J.C., 1969. Artefactos lticos en abrigos
rocosos de El Abra, Colombia. Informe preliminar. Revista Colombiana de
Antropologa 14, 10e46.

rrez, J., Caldero


 n, K., Villada, C., 2005. Evidencias arqueolo
gicas y
Correal, G., Gutie
megafauna en un salado del tardiglacial superior. Boletn de Arqueologa 20,
3e58.
Currie, L.A., 2004. The remarkable metrological history of radiocarbon dating [II].
Journal of Research of the National Institute of Standards Technology 109,
185e217.
Delgado, M.E., 2012a. Paleoecologa y poblamiento temprano del Noroccidente de
rica durante la transicio
n Pleistoceno/Holoceno (ca. 12,000e8000 an
~ os
Sur Ame
14
rica: modC AP). In: IV Simposio Internacional El hombre temprano en Ame
elos de poblamiento y aportes desde las territorialidades tropicales. Pereira,
Colombia.
Delgado, M.E., 2012b. Mid and late Holocene population changes at the Sabana de
 (northern South America) inferred from skeletal morphology and
Bogota
radiocarbon chronology. Quaternary International 256, 2e11.
Domnguez, C., 1988. El espacio. In: Historia de Colombia, vol. 1. Salvat, Bogot
a,
pp. 3e35.
Gamble, C.S., Davies, W., Pettit, P., Richards, M., 2004. Climate change and evolving
human diversity in Europe during the last glacial. Philosophical Transactions of
the Royal Society of London B 359, 243e254.
Geyh, M.A., De Maret, P., 1982. Research notes and application reports histogram
evaluation of 14C dates applied to the rst complete Iron Age sequence from
west central Africa. Archaeometry 24, 158e163.
Gillespie, R., Hedges, R.E.M., Wand, J.O., 1984. 14C dating of bone by accelerator mass
spectrometry. Journal of Archaeological Science 11, 165e170.
Gkiasta, M., Russell, T., Shennan, S., Steele, J., 2003. Neolithic transition in Europe:
the radiocarbon record revisited. Antiquity 77, 45e62.
Gnecco, C., 1994. The Pleistocene/Holocene Boundary in the Northern Andes: an
Archaeological Perspective (Unpublished PhD dissertation). Washington University, Department of Anthropology, Saint Louis.
n temprana de bosques tropicales de montan
~ a. Editorial
Gnecco, C., 2000. Ocupacio
Universidad del Cauca.
Gnecco, C., 2003. Against ecological reductionism: Late Pleistocene huntergatherers in the tropical forests of northern South America. Quaternary International 109e110, 13e21.
Gnecco, C., Aceituno, F., 2006. Early humanized landscapes in Northern South
America. In: Morrow, J.E., Gnecco, C. (Eds.), Paleoindian Archaeology:
a Hemispheric Perspective. University Press of Florida, Gainesville,
pp. 86e104.
Gnecco, C., Mora, S., 1997. Late Pleistocene/Early Holocene tropical forest occupa~ a Roja, Colombia. Antiquity 71, 683e690.
tions at San Isidro and Pen
Gowlett, J.A., 1987. The archaeology of radiocarbon accelerator dating. Journal of
World Prehistory 1, 127e170.
~ os antes del
Groot, A.M., 1992. Checua: Una secuencia cultural entre 8500 y 3000 an
n de Investigaciones Arqueolo
gicas Nacionales, Banco de la
presente. Fundacio
Repblica, Bogot
a.
rrez, M.A., 2004. An
micos en el 
Gutie
alisis tafono
area Interserrana (Provincia de
Buenos Aires) (Unpublished PhD dissertation). Facultad de Ciencias Naturales y
Museo, Universidad Nacional de La Plata, La Plata.
Hamilton, M.J., Buchanan, B., 2007. Spatial gradients in Clovis-age radiocarbon dates
across North America suggest rapid colonization from the north. Proceedings of
the National Academy of Sciences of the United States of America 104,
15625e15630.
Hedges, R.E.M., Van Klinken, G.J., 1992. A review of current approaches in the
pretreatment of bone for 14C dating by AMS. Radiocarbon 34, 279e291.
Herrera, L., Bray, W., Cardale, M., Botero, P., 1992. Nuevas fechas de radiocarbono
mico de la Cordillera Occidental de Colombia. In: Ortizpara el precera
Troncoso, O., van der Hammen, T. (Eds.), Archaeology and Environment in Latin
America. Instituut voor Pre- en Protohistorische Acheologie Albert Egges van
Giffen, Universiteit van Amsterdam, Amsterdam, pp. 145e163.
~ a, O., 2011. La historia muy antigua del municipio de
Herrera, L., Moreno, C., Pen
 gico del aeropuerto
Palestina (Caldas). Proyecto de rescate y monitoreo arqueolo
. Centro de Museos-Universidad de Caldas, Asociacio
 n Aeropuerto del
del cafe
 (2005e2011), Manizales.
Cafe
Hofstede, R., Segarra, P., Mena, P.V., 2003. Los P
aramos del Mundo. Global Peatland
Initiative/NC-IUCN/EcoCiencia, Quito.
Housley, R.A., Gamble, C.S., Street, M., Pettitt, P.B., 1997. Radiocarbon evidence for
the Lateglacial human recolonisation of Northern Europe. Proceedings of the
Prehistoric Society 63, 25e54.
Hughen, K.A., Overpeck, J.T., Lehman, S.J., Kashgarian, J.M., Southon, R.,
Petersson, L.C., 1998. A new 14C calibration data set for the last deglaciation
based on marine varves. Radiocarbon 40, 483e494.
Hurt, W., 1977. The edge-trimmed tool tradition of Northwest South America. In:
Cleland, C.E. (Ed.), For the Director. Research Essays in Honor of James B. Grifn,
Anthropology Papers, vol. 61. Museum of Anthropology, University of Michigan,
Ann Arbor, pp. 268e294.
Hurt, W., van der Hammen, T., Correal, G., 1977. The El Abra Rockshelters, Sabana de
, Colombia, South America. In: Occasional Papers and Monographs No. 2.
Bogota
Indiana University Museum, Bloomington.
INTEGRAL, 1997. Arqueologa de Rescate: Va Alterna de la Troncal de Occidente
Ro Campoalegre-Estadio Santa Rosa de Cabal (Informe Final). INTEGRAL
S.A., Ministerio de Transporte, Instituto Nacional de Vas, Medelln
(Unpublished).
ris, O., Ferna
ndez, E., Weninger, B., 2003. Radiocarbon evidence of the Middle to
Jo
Upper Paleolithic transition in southwestern Europe. Trabajos de Prehistoria 60,
15e38.

Please cite this article in press as: Delgado, M., et al., 14C data and the early colonization of Northwest South America: A critical assessment,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.09.011

10

M. Delgado et al. / Quaternary International xxx (2014) 1e10

pez, C., 1989. Evidencias paleoindias en el valle medio del ro Magdalena


Lo
 y Remedios, Antioquia). Boletn de
(Municipios de Puerto Berro, Yondo
Arqueologa 4 (2), 3e23.
n de Medio Ro Caqueta
Llanos, J.M., 1997. Artefactos de molienda en la regio
Amazona Colombiana. Boletn de Arqueologa 12 (2), 3e95.
pez, C., 2008. Landscape Development and the Evidence for Early Human
Lo
Occupation in the Inter-andean Tropical Lowlands of the Magdalena River,
Colombia. Syllaba Press, Miami.
Marchant, R., Behling, H., Berro, J., Cleef, A., Duivenvoorden, J., Helmens, K.,
Hooghiemstra, H., Kuhry, P., Melief, R., Schreve-Brinkman, E., van Geel, B., van
Reenen, G., van der Hammen, T., 2001. A reconstruction of Colombian biomes
derived from modern pollen data along an altitudinal gradient. Review of
Palaeobotany and Palynology 117, 79e92.
Marchant, R., Behling, H., Berro, J., Cleef, A., Duivenvoorden, J., Hooghiemstra, H.,
Kuhry, P., Kuhry, R., Melief, R., Schreve-Brinkman, E., van Geel, B., van der
Hammen, T., van Reenen, G., Wille, M., 2002. Pollen-based biome reconstructions for Colombia at 3000, 6000, 9000, 12000, 15000 and 18000 14C yr
ago: Late Quaternary tropical vegetation dynamics. Journal of Quaternary Science 17, 113e129.
Marchant, R., Boom, A., Behling, H., Hooghiemstra, H., Melief, B., van Geel, B., van
der Hammen, T., Wille, M., 2004. Colombian vegetation at the Last Glacial
Maximum: a comparison of model and pollen-based biome reconstructions.
Journal of Quaternary Science 19, 721e732.
Meese, D.A., Gow, A.J., Alley, R.B., Zielinski, G.A., Grootes, P.M., Ram, M., Taylor, K.C.,
Mayewski, P.A., Bolzan, J.F., 1997. The Greenland Ice Sheet Project 2 deptheage
scale: methods and results. Geophysical Research 102 (C12), 26,411e26,423.
 ska, D.J., Pazdur, A., 2004. Shape analysis of cumulative probability density
Michczyn
function of radiocarbon dates set in the study of climate change in the late
glacial and Holocene. Radiocarbon 46, 733e744.
Mora, G., Pratt, L.M., 2002. Carbon isotopic evidence from paleosols for mixed C3/C4
 basin, Colombia. Quaternary Science Review
vegetation in the Sabana de Bogota
21, 985e995.
Mora, S., Gnecco, C., 2003. Archaeological hunter-gatherers in tropical rainforests: a
view from Colombia. In: Mercader, J. (Ed.), Under the Canopy: the Archaeology
of Tropical Rain Forests. Rutgers University Press, New Brunswick, pp. 271e290.
Morcote, R.G., Cabrera, D., Mahecha, C.E., Franky, C., Cavellier, I., 1998. Las palmas
entre los grupos cazadores recolectores de la Amazonia colombiana. Caldasia
20, 57e74.
Newby, P., Bradley, J., Spiess, A., Shuman, B., Leduc, P., 2005. A Paleoindian response
to Younger Dryas climate change. Quaternary Science Reviews 24, 141e154.
Nieuwenhuis, C.J., 1998. Unattractive but effective: unretouched pointed akes as
projectile points? A closer look at the Abriense and Tequendamiense artefacts.
In: Plew, M.G. (Ed.), Explorations in American Archaeology. Essays in Honor of
Wesley R. Hurt. University Press of America, Lanham, pp. 133e163.
Nieuwenhuis, C.J., 2002. Traces on Tropical Tools: a Functional Study of Chert Artifacts from Preceramic Sites in Colombia (PhD thesis from Leiden University.
Archaeological Studies Leiden University No. 9). Faculty of Archaeology, University of Leiden, Leiden.
~o
n del ro Porce.
Otero, H., Santos, G., 2006. Las ocupaciones prehisp
anicas del can
 n, rescate y monitoreo arqueolo
gico. Proyecto hidroele
ctrico Porce
Prospeccio
III. Obras de infraestructura. Informe nal. Universidad de Antioquia-Empresas
Pblicas de Medelln, Medelln. (Unpublished).
~ oz, S.E., Gaweski, K., Viau, A.E., 2010. Prehistoric demography of
Peros, M.C., Mun
North America inferred from radiocarbon data. Journal of Archaeological Science 37, 656e664.
Pettit, P., Davies, W., Gamble, C.S., Richards, M.B., 2003. Radiocarbon chronology:
quantifying our condence beyond two half-lives. Journal of Archaeological
Science 30, 1685e1693.
Pinto, M., 2003. Galindo, un Sitio a Cielo Abierto de Cazadores-Recolectores en la
n de Investigaciones Arqueolo
 gicas Nacionales,
Sabana de Bogot
a. Fundacio
.
Banco de la Repblica, Bogota
Piperno, D., Pearsall, D., 1998. The Origins of Agriculture in the Lowland Neotropics.
Academic Press, San Diego.
Reichel-Dolmatoff, G., 1965. Colombia Ancient Peoples and Places. Thames and
Hudson, London.
Reichel-Dolmatoff, G., 1997. Arqueologa de Colombia: Un Texto Introductorio.
.
Presidencia de la Repblica, Bogota
gico en el a
rea de inuencia del
Restrepo, C.A., 2013. Informe de monitoreo arqueolo
.
Proyecto de Desarrollo Vial Armenia Pereira Manizales, Autopista de Cafe
 S. A., Pereira (Unpublished).
Instituto Nacional de Vas, Autopistas de Cafe
Rick, J.W., 1987. Dates as data: an examination of the Peruvian preceramic radiocarbon data. American Antiquity 52, 55e73.
Riede, F., 2009. Climate and demography in early prehistory: using calibrated 14C
dates as population proxies. Human Biology 81, 309e337.
~ os de Presencia Humana en el P
 n de
Rivera, S., 1991. Neusa 9000 An
aramo. Fundacio
gicas Nacionales, Banco de la Repblica, Bogota
.
Investigaciones Arqueolo
nico: Procesos Socioculturales
Rodrguez, C.A., 2002. El Valle del Cauca Prehispa
 ricas del Alto y Medio Cauca y la Costa
Antiguos en las Regiones Geohisto

Pacca Colombo-Ecuatoriana. Departamento de Historia, Facultad de Humanidades Universidad del Valle, Santiago de Cali.
rez, A., 2003. Recopilacio
 n y ana
lisis de las fechas Carbono-14 del Norte
Rubinos Pe
de Santa Cruz. In: Aguerre, A.M. (Ed.), Arqueologa y Paleoambiente en la
~ a Argentina, pp. 1e25 (Buenos Aires).
Patagonia Santacrucen
micos en el alto y medio ro Calima,
Salgado, H., 1988-1990. Asentamientos precera
Cordillera Occidental de Colombia. Cespedesia 16e17 (57e58), 139e162.
 gicas en la Cordillera Central RoncesvalSalgado, H., 1998. Exploraciones Arqueolo
n de Investigaciones Arqueolo
gicas Nacionales, Banco de
leseTolima. Fundacio
la Repblica, Bogot
a. Universidad del Tolima, Fondo Mixto de Cultura del Tolima, Bogot
a.
~ os de Ocupaciones Humanas en Envigado Antioquia. El
Santos, G., 2010. Diez Mil An
Sitio La Morena. Alcalda de Envigado, Envigado.
Stafford Jr., T.W., Jull, A.J.T., Brendel, K., Duhamel, R.C., Donahue, D., 1987. Study of
bone 14C dating accuracy at the University of Arizona NSF accelerator facility for
radioisotope analysis. Radiocarbon 29, 24e44.
Stafford Jr., T.W., 1990. Late Pleistocene megafauna extinctions and the Clovis culture: absolute ages based on accelerator 14C dating of skeletal remains. In:
Agenbroad, L.D., Mead, J.I., Nelson, L.W. (Eds.), Megafauna and Man: Discovery
of America's Heartland, Hot the Mammoth Site of Hot Springs, South Dakota,
Inc., Scientic Papers, vol. 1. Northern Arizona University, Flagstaff, pp. 118e122.
Steele, J., 2010. Radiocarbon dates as data: quantitative strategies to estimating
colonization front speed and event densities. Journal of Archaeological Science
37, 2017e2030.
Steele, J., Politis, G., 2009. AMS 14C dating of early human occupation of southern
South America. Journal of Archaeological Science 36, 419e429.
Surrovell, T., Brantingham, J.P., 2007. A note on the use of temporal frequency
distributions in the studies of prehistoric demography. Journal of Archaeological Science 34, 1868e1877.
n: Arqueologa
Tabares, D., Rojas, E., 2000. Aportes para una Historia en Construccio
de Rescate en la Doble Calzada Manizales-Pereira-Armenia. INVIAS-CISAN,
Bogot
a (Unpublished).
Taylor, R.E., 1987. Radiocarbon Dating: an Archaeological Perspective. Academic
Press, Orlando.
van der Hammen, T., 1974. The Pleistocene changes of vegetation and climate in
tropical South America. Journal of Biogeography 1, 3e26.
van der Hammen, T., 1978. Stratigraphy and environments of the Upper Quaternary
of the El Abra corridor rockshelters (Colombia). Palaeogeography, Palaeoclimatology, Palaeoecology 25, 111e162.
 n del Mastovan der Hammen, T., 1986. Cambios medioambientales y la extincio
donte en el norte de los Andes. Revista de Antropologa 2, 27e33.
n. Corporacio
n Colombiana
van der Hammen, T., 1992. Historia, Ecologa y Vegetacio
para la Amazonia Araraucara, Bogot
a.
nico en
van der Hammen, T., Correal, G., 2001. Mastodontes en un humedal Pleistoce
el valle del Magdalena (Colombia) con evidencias de la presencia del hombre en
el pleniglacial. Boletn de Arqueologa 16, 4e36.
lez, E., 1960. Upper Pleistocene and Holocene climate
van der Hammen, T., Gonza
 (Colombia, South America). Laidse
and vegetation of the Sabana de Bogota
Geologische Mededelingen 25, 261e315.
van der Hammen, T., Hooghiemstra, H., 1995. The El Abra stadial, a Younger Dryas
equivalent in Colombia. Quaternary Sciences Review 14, 841e851.
van Geel, B., van der Hammen, T., 1973. Upper Quaternary vegetational and climatic
sequence of the Fquene area (Eastern Cordillera, Colombia). Palaeogeography,
Palaeoclimatology, Palaeoecology 14, 9e92.
lez, M., Hooghiemstra, H., Metcalfe, S., Willie, M., Berro, J., 2006. Late Glacial and
Ve
Holocene environmental and climatic changes from a limnological transects
through Colombia, northern South America. Palaeogeography, Palaeoclimatology, Palaeoecology 234, 81e96.
Vermeersch, P.M., 2005. European population changes during Marine Isotope Stages
2 and 3. Quaternary International 137, 77e85.
Waterbolk, H.T., 1971. Working with radiocarbon dates. Proceedings of the Prehistoric Society 37, 15e33.
ris, O., 2008. A 14C age calibration curve for the last 60 ka: the
Weninger, B., Jo
Greenland-Hulu U/Th timescale and its impact on understanding the Middle to
Upper Paleolithic transition in Western Eurasia. Journal of Human Evolution 55,
772e781.
Williams, A.N., 2012. The use of summed radiocarbon probability distributions in
archaeology: a review of methods. Journal of Archaeological Science 39, 578e589.
Wilmshurst, J.M., Hunt, T., Lipo, C., Anderson, A., 2011. High precision radiocarbon
dating shows recent and rapid initial human colonization of East Polynesia.
Proceedings of the National Academy of Sciences of the United States of
America 108, 1815e1820.
Yuan, S., Wu, X., Gao, S., Wang, J., Cai, L., Liu, K., Li, K., Ma, H., 2000. Comparison of
different bone pretreatment methods for AMS 14C dating. Nuclear Instruments
and Methods in Physics Research Section B: Beam Interactions with Materials
and Atoms 172, 424e427.

Please cite this article in press as: Delgado, M., et al., 14C data and the early colonization of Northwest South America: A critical assessment,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.09.011

Quaternary International xxx (2014) 1e13

Contents lists available at ScienceDirect

Quaternary International
journal homepage: www.elsevier.com/locate/quaint

Teosinte before domestication: Experimental study of growth and


phenotypic variability in Late Pleistocene and early Holocene
environments
Dolores R. Piperno a, b, *, Irene Holst b, Klaus Winter b, Owen McMillan b
a
b

Department of Anthropology, Program in Human Ecology and Archaeobiology, Smithsonian National Museum of Natural History, Washington, DC, USA
Smithsonian Tropical Research Institute, Panama

a r t i c l e i n f o

a b s t r a c t

Article history:
Available online xxx

Agriculture arose during a period of profound global climatic and ecological change following the end of
the Pleistocene. Yet, the role of phenotypic plasticity e an organisms ability to change its phenotype in
response to the environment e and environmental inuences in the dramatic phenotypic transformations that occurred during plant domestication are poorly understood. Another factor possibly
inuential in agricultural origins, the productivity of crop plant wild progenitors in Late Pleistocene vs.
Holocene environments, has received increasing attention recently and merits further investigation. In
this study, we examined phenotypic characteristics and productivity (biomass, seed yield) in the wild
progenitor of maize, the teosinte Zea mays ssp. parviglumis H.H. Iltis & Doebley, when it was rst
exploited and cultivated by growing it in atmospheric CO2 concentrations and temperatures characteristic of the late-glacial and early Holocene periods. Plants responded with a number of attributes uncharacteristic of teosinte in todays environments, including maize-type traits in vegetative architecture,
inorescence sexuality, and seed maturation. Teosinte productivity was signicantly lower in late-glacial
compared with early Holocene and modern environments. Our evidence indicates that: a) ancestral
biological characteristics of crop plant progenitors arent always predicted from living examples, b) some
important maize phenotypic traits were present at initial human exploitation and selection, and c)
Pleistocene plant productivity should be considered a signicant factor in the chronology of food production origins.
2014 Elsevier Ltd and INQUA. All rights reserved.

1. Introduction

domesticated, Late Pleistocene (c. 21e11 ka) temperature and


annual precipitation were also signicantly lower than early postglacial levels by approximately 5e7 C and 20e40%, respectively
(Piperno et al., 2007; Hodell et al., 2008; Bush et al., 2009; CorreaMetrio et al., 2012). Temperature and atmospheric CO2 were still
lower than today by a few degrees Centigrade and about 130 ppmv
at the beginning of the Holocene (Ahn et al., 2004; Correa-Metrio
et al., 2012).
For a number of crop and wild progenitor species, recent
research has considerably elucidated the genetic mechanisms that
underwrote their phenotypic transformations from wild to
domesticated status in those climate eras. Previously emphasized
conventional assumptions for morphological change (e.g., that it
was driven by human selection for rare mutants of single genes that
were deleterious in wild plants and favorable in eld environments,
or selection for new, advantageous mutations that appeared postcultivation) have for some domestication traits been supplanted
by more complex processes. They include epistasis (when the
phenotypic product of one gene depends on its interactions with

The development of agricultural societies made possible by


plant and animal domestication was one of the most transformative
events in human and ecological history. Agricultural beginnings can
be traced around the world to 12,000e10,000 years ago (ka) during
a time of profound global environmental change as the Pleistocene
was ending and transitioning to the Holocene (e.g., Price and BarYosef, 2011). From at least 40,000e12,000 ka, atmospheric CO2
levels worldwide were as much as one-third lower (c. 180e
235 ppmv) than early (12e10 ka) post-glacial levels (c. 265e
270 ppmv) (Ahn et al., 2004). In many areas of the world, including
the New World tropics where maize and other major crops were

* Corresponding author. Department of Anthropology, Program in Human Ecology and Archaeobiology, Smithsonian National Museum of Natural History,
Constitution Ave., Washington, DC, USA.
E-mail address: pipernod@si.edu (D.R. Piperno).
1040-6182/$ e see front matter 2014 Elsevier Ltd and INQUA. All rights reserved.
http://dx.doi.org/10.1016/j.quaint.2013.12.049

Please cite this article in press as: Piperno, D.R., et al., Teosinte before domestication: Experimental study of growth and phenotypic variability in
Late Pleistocene and early Holocene environments, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2013.12.049

D.R. Piperno et al. / Quaternary International xxx (2014) 1e13

other genes in the same plant), non-Mendelian inheritance, and


changes in gene expression (such as how, when, and to what extent
existing genes are expressed through changes in the amount of
mRNA during transcription) (e.g., Nesbitt and Tanksley, 2002;
Doebley, 2004; Studer and Doebley, 2011; Studer et al., 2011;
Swanson-Wagner et al., 2012; Olsen and Wendel, 2013). Furthermore, in an increasing number of wild progenitors, pre-existing,
non-deleterious genetic variation also called standing and
cryptic genetic variation is being documented for major domestication traits, indicating genetic attributes for the traits were
commonly available to the rst plant cultivators (Lauter and
Doebley, 2002; Nesbitt and Tanksley, 2002; Clark et al., 2004;
Studer et al., 2011).
It is now also well-established through other research that the
environment may inuence some of these geneticephenotypic
relationships by directly inducing (triggering) phenotypic variability (e.g., West-Eberhard, 2003; Gilbert and Epel, 2009; Beldade
et al., 2011; Palmer et al., 2012). Phenotypic plasticity e when a
single genotype exhibits multiple phenotypes in response to
environmental variability and change e is well-documented in
plants, and there is increasing evidence that phenotypes generated
in this manner can be both adaptive and inherited (e.g., WestEberhard, 2003; Moczek, 2007; Gilbert and Epel, 2009; Nicotra
et al., 2010; Sultan, 2010; Beldade et al., 2011; Moczek et al.,
2011; Palmer et al., 2012). However, how natural- and humanmediated environmental change may have inuenced the dramatic phenotypic transitions undergone by domesticated plants is
a neglected area of domestication research.
It, therefore, becomes of considerable interest to ask if, during
the late-glacial and early Holocene periods (c. 16e10 ka) when
people rst encountered, exploited, and cultivated many of the wild
progenitors, the plants differed from modern wild populations,
inuencing crop plant evolution in ways that have been little
considered. The last hunters and gatherers and rst farmers
worked with the phenotypes they saw, and it can be imagined they
were attuned to and interested in the phenotypic variability they
encountered on natural and cultivated landscapes. Unfortunately,
chronologically-coarse and often geographically-uneven archaeobotanical records do not adequately capture the range of phenotypic attributes that early cultivators experimented with. Moreover,
the macrofossils (seeds, fruits, stems) that can best inform this
question are often poorly preserved and have as yet to be recovered
from Late Pleistocene and early Holocene records for many wild
progenitors and earliest cultivars, including Zea (e.g., Piperno et al.,
2009; Piperno, 2011). Thus, modern representatives of crop plant
ancestors constitute the basis for much of the morphological and
genetic study of proto-domesticates and their wild ancestors.
A related question of increasing interest concerns the productivity of wild progenitors of crops in Pleistocene environments
before farming began. Plant growth is highly CO2 dependent and
low Pleistocene atmospheric CO2 appears to have been a signicant
stress factor on many species, lowering their productivity because
of its constraints on photosynthesis and water-use efciency
(Dippery et al., 1995; Sage, 1995; Cunniff et al., 2010; Gerhart and
Ward, 2010; Cowling, 2011). It is increasingly seen as having been
a major limiting factor for a pre-Holocene origin of agriculture
(Sage, 1995; Richerson et al., 2001; Cunniff et al., 2008, 2010).
Recent experimental studies on a number of C3 and C4 cereal crop
progenitors, including wild maize e the teosinte Zea mays ssp.
parviglumis H.H. Iltis & Doebley e found decreased biomass and/or
photosynthetic activity when plants were grown in Pleistocene
compared with Holocene CO2 levels (Cunniff et al., 2008, 2010).
Additional constraints on productivity during the Pleistocene may
have included depressed temperatures and precipitation, but the
impact of these factors has not been studied.

In order to experimentally investigate the phenotypic and


productivity variability encountered by hunters-gatherers and
proto-farmers, we grew maizes wild ancestor (hereafter ssp. parviglumis) in glass houses in which temperature and CO2 levels were
adjusted to those documented in Mesoamerica for the late-glacial
segment (c. 16e11 ka) of the Late Pleistocene and the beginning
of the Holocene (c. 11e10 ka) (e.g., Ahn et al., 2004; Piperno et al.,
2007; Bush et al., 2009; Correa-Metrio et al., 2012). Our results
address the possible multifactorial roles of phenotypic variability
and plasticity, plant productivity, and environmental change in
constructing theories for, investigating, and understanding maize
domestication and perhaps crop plant origins more generally.
2. Teosinte, phenotypic plasticity, and environmental change:
background and previous relevant research
Maizes wild ancestor appears to be a particularly suitable candidate with which to begin evaluating the links between phenotypic
and environmental variability in domestication research. Archaeological and genetic evidence indicates that maize was domesticated
in tropical southwestern Mexico, probably in the central Balsas River
Valley region of Michoacn and Guerrero states, by 9000 BP
(Matsuoka et al., 2002; Piperno et al., 2009; Ranere et al., 2009; Van
Heerwaarden et al., 2011). Although not yet documented with
archaeological data, the likelihood is that teosinte cultivation began at
least 1000 years earlier (Wang et al., 2005), shortly after the Pleistocene ended and the climate and vegetation were in a state of
considerable transformation. There are profound morphological differences between teosinte and maize in vegetative architecture and
inorescence sexuality that are the most dramatic known of any crop/
progenitor pair, a factor that led to a century-long debate about
maizes ancestry (discussed in Doebley, 2004). The differences are
known to be in part underwritten by the gene teosinte branched1 (tb1)
through a gene expression change (change in the amount of mRNA) it
mediates during early plant development (Doebley et al., 1995;
Hubbard et al., 2002; Studer et al., 2011). Given this observation and
reasons discussed below, a type of plasticity called developmental
plasticity may have been relevant to teosinte and its domestication.
Developmental plasticity is the inherent capacity of organisms
to rapidly produce phenotypic change through one of several
available pre-adult developmental pathways in direct response to
environmental perturbations and stress factors (e.g., WestEberhard, 2003; Moczek, 2007; Fusco and Minelli, 2010; Gilbert
and Epel, 2009; Phennig et al., 2010; Sultan, 2010; Beldade et al.,
2011; Moczek et al., 2011). This capacity should be particularly
important in plants, which cannot simply get up and move to places
more to their liking when physical and biotic conditions become
less favorable. Developmental plasticity is the integration of
evolutionary developmental biology (evo-devo) with environmental inuences in the determination of what causes phenotypic
change. The environment can include factors emanating externally
or from within the internal conditions of organisms. Developmental plasticity has become closely allied with the new eld of
ecological developmental biology (eco-devo or eco-evo-devo),
which places particular emphasis on external environmental inuences on phenotypes that broadly range from the physical
environment to competitors and predators, and eld rather than
laboratory research (Gilbert and Epel, 2009; Sultan, 2010; Beldade
et al., 2011). Developmental plasticity often gives rise to new
phenotypes through changes in gene expression, which is known to
be highly responsive to environmental perturbation (e.g., Gilbert
and Epel, 2009; Beldade et al., 2011). New phenotypic variation
can then be rapidly introduced without a corresponding genetic
change (e.g., without the appearance or spread of a new mutation),
and it can spread in populations if the inducing environment is

Please cite this article in press as: Piperno, D.R., et al., Teosinte before domestication: Experimental study of growth and phenotypic variability in
Late Pleistocene and early Holocene environments, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2013.12.049

D.R. Piperno et al. / Quaternary International xxx (2014) 1e13

Fig. 1. The differences between teosinte and maize in branching architecture and inorescence sexuality. Teosinte has many long primary lateral branches terminated by tassels,
and secondary lateral branching. The female ears are located on the secondary lateral branches. Modern maize has a single main stem with a solitary tassel terminating it. There are
few, very short primary lateral branches, and no secondary branching. The cobs are located at the ends of the short primary lateral branches in the positions occupied by tassels in
teosinte. Credit: Nicolle Rager Fuller, National Science Foundation.

maintained over multiple generations. These points will be elaborated on further below.
Importantly, a developmentally plastic response in vegetative
architecture and inorescence sexuality takes place in teosinte
today that adapts plants to their local environments. In good
growing conditions (adequate sunlight, deep soils), the plant is tall
e 2e3 m-high e with many long lateral branches tipped by tassels
and secondary branches bearing female ears with a few small seeds
(Fig. 1, left). These are the vegetative and oral characteristics
normally associated with maizes wild ancestor both today and in
the past. However, stressful or less optimal habitats today (shade,
shallow soils, low moisture, crowding) induce a gene expression
change that causes suppression of branch elongation during
growth (Doebley et al., 1995; Hubbard et al., 2002; Whipple et al.,
2011). The result is plants with maize-like attributes; namely, a
few, dramatically shortened lateral branches tipped by female ears
instead of tassels and a single tassel terminating the main stem
(Fig. 1, right). These plants can also be very short (knee-high). It is
reasonable to believe that teosinte plasticity today in sub-optimal
growing conditions may be non-specic responses to a variety of
environmental stresses/cues, and that past conditions such as low
CO2 and temperature may have been among them. In all environments, however, the domesticated maize exhibits a few, very short
lateral branches terminating in large cobs instead of tassels, and has
a solitary tassel at the top of a single main stem (Fig. 1, right). As in
teosinte, these transformations are in part mediated by the gene
expression and developmental changes discussed above (Doebley
et al., 1995; Hubbard et al., 2002). Maize domestication then
involved a loss of plasticity in these traits because maize vegetative
architecture and oral sexuality are constitutively expressed
regardless of the environment.
With a theory of maize evolution called Catastrophic Sexual
Transmutation, Iltis (1983, 1987) rst drew major attention to how
the vegetative (also called branching) architecture and inorescence traits of Zea discussed above were determined by mechanisms set in motion during early plant development. He
emphasized how environmental factors, including cold growing
seasons, were likely triggers, potentially producing a rapid
phenotypic transformation from teosinte- to maize-type branching

and inorescence sexuality without human involvement. A part of


Iltis theory that proposed the maize ear was derived from a
feminized teosinte tassel is probably incorrect, as subsequent
studies showed that when lateral branches were shortened and
female ears were translocated to the maize cob position at the ends
of the branches, they were of normal teosinte type with hard
cupulate fruitcases and didnt possess characteristics, such as soft
glumes, that would have been derived from tassels if the latter were
involved (Doebley et al., 1995). Nonetheless, Iltis focus on environmental inuences, plasticity, pre-existing genetic variation, and
the rapid, macro-evolutionary nature of such non-mutational
phenotypic change was ahead of its time in domestication
research and clearly warrants further attention.
Given both mechanisms for a direct environmental determination of phenotypic change, and a hypothesis for the involvement of
such a mechanism in maize domestication, an important initial
question for domestication studies becomes, what is the sensitivity of teosinte and other crop progenitors to past environmental
inuences?
3. Material and methods
3.1. Experiment design
Data presented here are from grow-outs of ssp. parviglumis from
four different natural Mexican populations and two lines of inbred
teosinte during its natural growing period from July to December in
two naturally-lighted glass environmental chambers housed at the
Gamboa eld station at the Smithsonian Tropical Research Institute
in Panama (Table 1). One chamber was adjusted to either late-glacial
or early Holocene temperature, and CO2 sub-ambient levels (Ahn
et al., 2004; Piperno et al., 2007; Hodell et al., 2008; Bush et al.,
2009; Correa-Metrio et al., 2012). The other was at modern CO2
levels and temperatures characteristic of ssp. parviglumis environments today (Table 2). We repeated the experiment in late-glacial
conditions three times from 2009 to 2011, and in 2012 conducted
the experiment in early Holocene conditions. In all years, plants
were germinated from seed in the chambers so that all pre-adult
development took place under the conditions being tested.

Please cite this article in press as: Piperno, D.R., et al., Teosinte before domestication: Experimental study of growth and phenotypic variability in
Late Pleistocene and early Holocene environments, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2013.12.049

D.R. Piperno et al. / Quaternary International xxx (2014) 1e13

Table 1
Sources of the teosinte seeds.
Accession

Origin

Plant name

Elevation asl

Population 1

PI 384062

B-K4

1350 m

Population 2

PI 384063

B-K7

1300 m

Population 3

PI 384071

Wilkes 10

1100 m

Population 4

PI 566692

Collected by J. Sanchez and G. Wilkes

850 m

RIMPA 0064

Ames 28408
06ncao01 SD
Ames 28409
06ncao01 SD

Mexico, Guerrero State


East side of highway, 1 mi s of Palo Blanco,
Latitude: 17 25 min 0 s N (17.41666667),
Longitude: 99 30 min 0 s W ( 99.5)
Mexico, Mexico State
West side of road, 4 km s of Valle de Bravo,
Latitude: 18 50 min N (18.83333333),
Longitude: 100 10 min 0 s W ( 100.16666667)
Mexico, Guerrero State
IgualaeArcelia Rd. 103 km from Iguala
Latitude: 18 20 min N (18.33333333),
Longitude: 100 19 min 0 s W ( 100.31666667)
Mexico, Michoacn State
Km 43 Rd. Zitacuaro to Tuzantla
19 4 min 0 s N (19.06666667),
Longitude: 100 25 min 0 s W ( 100.41666667)
Pedigree Beadle & Kato: Site 4

RIMPA 0065

Pedigree Wilkes: Site 6

Table 2
Temperatures in the chambers (C ).
Year
2009

2010

2011

2012

Mean
Mean
Mean
Mean
Mean
Mean
Mean
Mean
Mean

Min
Max
Min
Max
Min
Max

Mean
Mean Min
Mean Max

Range

Range

Late Glacial Chamber


21.3
20.3e23.5
19.6
17.9e20.2
27.8
21.4e33.6
22.5
20.6e29.7
18.7
17.3e24.5
30.3
24.4e44.7
20.1
18.1e22.7
15.5
14.5e16.3
30.2
21.8e37.2
Early Holocene Chamber
23
17.2e41.9
15.9
9.3e24.4
34.0
22.2e45.7

Modern Control Chamber


25.5
24.4e28.0
24.0
22.6e25.8
31.3
25.7e36.3
26.1
24.1e29.3
24.2
23.0e24.9
31.4
24.8e45.1
23.2
21.5e31.2
19.8
18.5e22.7
32.7
25.2e48.8
Modern Control Chamber
24.8
21.43e34.16
17.9
16.0e26.5
36.2
23.85e53.12

3.2. Plant descriptions


Non-inbred teosinte seeds were provided by the USDA North
Central Regional Plant Introduction Station located in Ames,
Iowa. They are from previous collections made by various investigators in four different localities in the states of Guerrero,
Michoacn, and Mexico and belong to four discrete populations
(Table 1). In the rst grow-out in 2009, three different populations, labeled Nos. 1e3, were planted; a fourth population was
added in 2010 and 2011, and included in the 2012 study. Collected
from altitudes between 850 and 1350 masl, they provide a good
representation of the elevational range of ssp. parviglumis today.
There was no evidence for maize introgression in the source
plants and none were collected from around existing maize
elds. One of the populations, from 1 mile south of Palo Blanco in
Guerrero, rarely hybridizes with maize and is thought to be the
least similar to maize of the annual teosintes (Wilkes, 1977).
Inbred teosinte seeds (RIMPA 0064 and 0065, listed as ssp. parglumis Ames 28408 and 28409 in the USDA Grin data base), that
became available to us for the 2012 study were supplied by Jeffrey Ross-Ibarra.
3.3. Growing conditions
The two naturally lighted glass-houses each have a c. 27 m3
internal volume (Fig. 2). Light intensity in the houses is about 80% of

the natural solar irradiation. Split air-conditioning units provided


temperature control. In the sub-ambient chambers, CO2 was lowered using a CO2 scrubber consisting of an acrylic column lled
with soda-lime, and air-pump and dust-bag of an industrial vacuum cleaner. Operation of the CO2 scrubber was under feedback
control via a Vaisala CO2 analyzer and a Campbell datalogger. CO2
values were logged every minute and overall mean values were
computed for the full growth period. The CO2 analyzer had a resolution of 10 ppmv and average daily CO2 concentration generally
varied between 200 and 220 ppmv. In the chamber adjusted to lateglacial conditions, average CO2 and temperature levels over the
three growth periods were from 203 to 216 ppmv and 20 to 22.5  C,
respectively (average CO2 in 2009 was 216 ppmv; in 2010,
203 ppmv; in 2011, 215 ppmv). Table 2 contains average diurnal,
maximum (daytime), and minimum (night), chamber temperatures. Modern ambient CO2 in the experiment area ranged from
360 ppmv in 2009 to 410 ppmv in 2012.
The 2012 grow-out was conducted at targeted, sub-ambient early
Holocene temperature and CO2 levels (c. 23  C, 260e265 ppm)
(Table 2). Daily average CO2 toward the end of the grow-out became
lower due to lower morning values, with the result that at the point
when every plant in the early Holocene chamber was mature, CO2
for the grow-out averaged to 252 ppmv. Average CO2 was 258 ppm
by the time seven of the 13 plants, including the two maize-like
phenotypes grown from MLP seeds (discussion below), were
mature; and 255 ppmv when nine plants, including one of the inbred
MLPs (discussion below), had matured. The remaining plants grew a
small amount in stature during their last few weeks of growth.
Therefore, phenotypic and yield results, including the induced
maize-like phenotypes, reect CO2 levels very close to the targeted
value.
3.4. Sample size and sampling
In the 2009e2011 grow-outs, from nine to 12 plants were grown
in each chamber in ve-gallon pots in natural top soil from a local
orchard without fertilizers. Three pots for each population were
planted with three seeds each to allow for non-germinating seeds;
after germination one plant per pot was allowed to grow to
maturity. In the 2012 grow-out, the sub-ambient and Modern
Control Chambers had the following: four plants grown from seeds
of maize-like phenotypes that were induced in the sub-ambient

Please cite this article in press as: Piperno, D.R., et al., Teosinte before domestication: Experimental study of growth and phenotypic variability in
Late Pleistocene and early Holocene environments, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2013.12.049

D.R. Piperno et al. / Quaternary International xxx (2014) 1e13

Fig. 2. The Sub-ambient chamber set at late-glacial conditions with teosinte.

chambers from 2009 to 2011 (plants 3C-2009, 4C-2010, 1A-2011,


2B-2011); one plant each from the two different inbreds, and two
replicate plants (sub-ambient) or one plant (modern control) from
the four different populations of the founder seed collections (in
one sub-ambient pot with seeds from founder population 2,
germination didnt occur). Pots were watered from two to four
times per week. During the ve to six month growth period various
aspects of plant development were recorded at least once a week,
such as height, branch length and number, and inorescence

characteristics. After plants matured they were harvested and


transported to the lab where additional descriptions and measurements were made (e.g., seed number/size/weight; biomass). In
order to take into account differences in plant height a ratio of
branch length to plant height was used to determine nal branch
length at maturity, called here relative branch length. Biomass was
measured after plants were air dried on the sum of four component
parts; seeds, leaves, other vegetative parts (stems, leaf sheaths, ear
bracts) and roots.

Fig. 3. A. A Maize-like phenotype plant from the Late Glacial Chamber. Like maize, it has a single tassel that terminates the main stem, female ears at the main stem (arrows) that
terminate a few, very short lateral branches, and no secondary branching. The inset at the upper right is a close-up of one of the female ears. B. Teosinte in the Modern Control
Chamber. Like in modern natural populations, it has many long, primary lateral branches (example, upper white arrow) terminated by tassels (black arrow). Female ears, not yet
developed, would be on secondary lateral branches at the location of the two bottom white arrows.

Please cite this article in press as: Piperno, D.R., et al., Teosinte before domestication: Experimental study of growth and phenotypic variability in
Late Pleistocene and early Holocene environments, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2013.12.049

D.R. Piperno et al. / Quaternary International xxx (2014) 1e13

4. Results
4.1. Phenotypic changes in branching architecture and inorescence
sexuality in late-glacial environments
We observed major phenotypic differences between the plants
from the sub-ambient chamber adjusted to late-glacial conditions
and modern control chamber. In every grow-out some of the plants
in the chamber with late-glacial conditions (hereafter, referred to as
the LGC) were complete maize-like phenotypes in branching architecture and inorescence sexuality; like maize, they had a few,
very short (non-measure-able) lateral branches tipped by female
ears instead of tassels, that were attached directly to a single main
stem tipped by a tassel (Fig. 3A) (hereafter, these plants with maizetype branching and inorescence sexuality traits are referred to as
maize-like phenotypes or MLPs). A total of six plants out of 33
from all the grow-outs combined representing every population
studied had these characteristics (one plant from population 3 in
2009; one from pop. 4 in 2010; and one each from pops. 1 and 2,
plus two from pop. 3 in 2011). Female ears of these plants, although
translocated to the ends of the primary branches at the main stems
and thus in positions homologous to where tassels are located in
teosinte and cobs in maize, were of normal teosinte type. They were
typically composed of 5e12 hard, disarticulated, cupulate fruitcases
(consisting of a kernel enveloped by a glume and rachid) subtended by vegetative bracts (Figs. 3 and 4). In contrast, the maizelike phenotypes did not occur in any grow-out in the modern
control chamber (hereafter, MCC), where plants were much like
those seen in natural environments today; they had many long
lateral branches terminated by tassels and secondary branches
terminated by female ears (Fig. 3B).
The maize-like phenotypes also exhibited a seed maturation
strategy characteristic of maize, with most of their seeds maturing
at the same time. In contrast, in other plants in the LGC and MCC, as
in the wild, seeds matured sequentially from the tips of the
branches to the base over a period of about two months, requiring
several harvest trips to collect them before they began to fall off
the plant soon after maturation. Overall in the three grow-outs, LGC
plants had fewer (Table 3) and shorter lateral branches (mean
relative branch length of 0.37  0.03 in the LGC vs. 0.42  0.03 in
the MCC). Furthermore, comparing branch length with the sexuality of the ower terminating the branches, it can be seen the two
traits are highly correlated. Average relative branch length when
terminated by tassels tended to be longer than when terminated by
mixed maleefemale inorescences (mixed MeF) (explained
below), and were signicantly longer than those terminated by

Fig. 4. Seeds enclosed in hard cupulate fruitcases (glume rachid) from a maize-like
phenotype plant that were removed from the vegetative bracts.

females (Fig. 5). Branches tipped by mixed maleefemale inorescences were signicantly longer than branches terminated by
female ears, which were always on the shortest branches on the
plant. The data provide signicant support to arguments that
branch length strongly inuences oral sexuality (Iltis, 1983).
Another nding relating to oral sexuality is that a nearly complete
feminization of the primary lateral branches occurred in the LGC,
where inorescences were almost always either completely feminine or mixed MeF (Fig. 6). Only one lateral branch out of a total of
128 in the LGC had a tassel. Therefore, LGC plants exhibited signicant similarities to maize regardless of whether they became
the complete, maize-like phenotypes in branching and inorescence sexuality.

Table 3
Mean branch and node number.
LGC or EHC

Mean 2009
Mean 2010
Mean 2011
Mean 09e11
Mean 2012

MCC

Branches

Nodes

Branches

Nodes

4
4
4
4
5

5
6
5
5
7

5
7
6
6
10

9
8
11
9
13

 1.2
 0.9
 1.3
0.2
 1.8

 1.1
 0.9
 1.3
0.2
 1.9

 1.2
 1.7
 2.1
0.3
 2.1

 2.3
 1.6
 1.5
0.6
 5.7

Means  SD, Mean 09e11  SEM. For branch No. in MCC vs. LGC 09e11 P < 0.001;
For node No. in MCC vs. LGC 09e11 P < 0.001; For branch No. in LGC 09e11 vs. EHC
2012 P 0.127 (not signicant); For node No. in LGC 09e11 vs. EHC P 0.006; For
branch No. in EHC vs. MCC 09e11 P 0.042; For node No. in EHC vs. MCC 09e11
P 0.042. Statistical signicance tested by the ManneWhitney rank sum test.
The LGC is the chamber adjusted to late-glacial conditions in years 2009e11 and the
EHC is the chamber adjusted to early Holocene conditions in 2012.

Tillering, the production of branch-like organs at the bottom of


plants, also differs signicantly in modern teosinte and maize (e.g.,
Hubbard et al., 2002; Whipple et al., 2011). Teosinte may typically
tiller, particularly when grown at temperate latitudes, whereas

Fig. 5. A box plot of relative branch length vs. inorescence sexuality of primary lateral
branches for data from 2009 to 2011 combined for the Late Glacial Chamber and
Modern Control Chamber. Notes: for MCC, P < 0.05 for mixed vs. female, P > 0.05 for
mixed vs. male, P < 0.05 for male vs. female. For LGC, P < 0.05 for mixed vs. female,
P > 0.05 for mixed vs. male, P < 0.05 for male vs. female. Statistical signicance
analyzed by the KruskaleWallis one way analysis of variance on ranks and the Dunns
method for all pairwise multiple comparison procedures. No tassels occurred in the
LGC in 2009 and 2011 and one occurred in the LGC in 2010.

Please cite this article in press as: Piperno, D.R., et al., Teosinte before domestication: Experimental study of growth and phenotypic variability in
Late Pleistocene and early Holocene environments, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2013.12.049

D.R. Piperno et al. / Quaternary International xxx (2014) 1e13

branches of intermediate length. They have been briey noted by


other researches in F2 populations of maize/teosinte hybrids and in
plants in the wild growing under sub-optimal conditions where
branch shortening has occurred (Doebley et al., 1990). Also of interest is that mixed MeF seeds often were not enclosed in vegetative bracts and were completely exposed, unlike normal female
ears (Fig. 7, compare with Fig. 3A).
4.2. Other phenotypic characteristics in the late-glacial
environment
Another major difference between the chambers was the short
stature of the plants in the LGC, which on average were less than
half the height of the MCC plants (Mean: 94  4.6 cm in the LGC vs.
226  10.4 cm in the MCC) (Fig. 8). There were no apparent differences in germination ability as most seeds in the LGC (Mean:
84%) and MCC (Mean: 82%) readily germinated in every grow-out.
4.3. Phenotypic plasticity and variability in the early Holocene
environment

Fig. 6. The percentages of inorescences of all sexualities terminating primary lateral


branches in the sub-ambient (top) and Modern Control Chambers (bottom) for the
years 2009e2012. The LGC is the chamber adjusted to late-glacial conditions in years
2009e11 and the EHC is the chamber adjusted to early Holocene conditions in 2012.

maize less commonly does so. This attribute was highly variable in
the grow-outs. Many plants in both chambers produced tillers in
2010, few in either chamber had them the other two years. A
number of factors are thought to control tillering in Zea and other
grasses, including the identied domestication genes associated
with branching and inorescence sexuality (Hubbard et al., 2002;
Doust, 2007; Whipple et al., 2011). One out of the six maize-like
phenotypes had tillers, suggesting an association between genes
thought to control tillering and environmental cues, as would be
expected (Doust, 2007; Whipple et al., 2011), although sample size
may not be large enough to make robust correlations. Feminization
in the LGC was also true of many tillers, all but two of which were
male in the MCC.
The mixed maleefemale inorescences (mixed MeF), which
commonly terminated lateral branches in the LGC and also
occurred in the MCC, contained seeds positioned proximally to the
lateral branch attached to male owers terminating the branch
(Fig. 7). In both chambers they are most typically present on

Teosinte foragers probably transitioned into persistent cultivators around the beginning of the Holocene, by c. 10.5e10 ka, when
Greenland ice core data and paleoenvironmental reconstructions
from Mesoamerican lakes indicate that atmospheric CO2 and
annual temperature were still depressed by more than 100 ppmv
(at c. 260e265 ppmv) and about two degrees (at c. 23  C) compared
with conditions experienced by modern teosinte (Ahn et al., 2004;
Wang et al., 2005; Piperno et al., 2007, 2009; Bush et al., 2009;
Correa-Metrio et al., 2012). We conducted a grow-out at these
and modern control conditions using the following: seeds from
four maize-like phenotype plants that were induced in the LGC
from 2009 to 2011; founder seeds from the four original ssp. parviglumis populations; and seeds from two inbred lines of teosinte
(plants self-pollinated for multiple generations so that they have
the same genotype) (see Methods: Sample size and sampling for
more details). Most seeds in both chambers germinated readily
again. In the chamber adjusted to early Holocene conditions
(hereafter, EHC), two out of four plants grown from maize-like
phenotype seeds were MLPs with uniform seed maturation again.
Branches of the other two, while longer than in the MLPs,
continued to be feminized; i.e., terminated by either completely
female ears or mixed maleefemale owers. Three out of four of
these plants, including the two MLPs, were short-statured (52e
79 cm-high; the other reached 152 cm). Both lines of inbreds also
became short-statured (83e84 cm-high), maize-like phenotypes in
the EHC. Responses of plants grown in the EHC from founder seeds
were more variable. Five of the seven had lateral branches terminated by female ears or mixed maleefemale owers and were
short-statured (55e129 cm-high), as was common in late-glacial
conditions in 2009e2011. The other two had long lateral
branches terminated by tassels, as normal in modern conditions;
one of these was tall (181 cm) (Fig. 8 for mean height data for all
plants). None of the founder seed plants became MLPs.
In summary, in early Holocene conditions, maize-like phenotypes were both reproduced in a second generation from rst
generation, induced MLP plants, and induced from seeds of longbranched, tall teosinte. A few plants from founder seeds responded to the conditions with attributes typical of modern-day plants;
however, most continued to differ from modern plants in the same
vegetative, inorescence, and stature traits shown to differ in lateglacial environments (Figs. 6, 8 and 9; mean relative branch length
for all plants in the EHC was 0.39  0.06). In contrast, no maize-like
phenotypes occurred in the modern control chamber. Plants were
tall with many branches or branch nodes, and almost all owers

Please cite this article in press as: Piperno, D.R., et al., Teosinte before domestication: Experimental study of growth and phenotypic variability in
Late Pleistocene and early Holocene environments, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2013.12.049

D.R. Piperno et al. / Quaternary International xxx (2014) 1e13

terminating lateral branches were tassels (Figs. 6, 8 and 9) (Table 3).


Many plants still had under-developed branches without owers
and branch nodes without branches when the grow-out was halted
after six months. It is unclear what caused these features.
4.4. Plant productivity in late-glacial and early Holocene
environments
Previous experimental research on a range of C3 and C4 crop
plant progenitors including wild barley (Hordeum spontaneum K.
Koch), foxtail millet (Setaria viridis (L.) P. Beauv.), and ssp. parviglumis showed that photosynthesis and biomass increased when
CO2 was raised from 180 ppmv, representing its lowest level of the
last glacial period at c. 20 ka, to early Holocene levels of 270 ppmv
(Cunniff et al., 2008, 2010). Teosinte seed yield and weight were not
reported. Our results also indicate that teosinte productivity is
signicantly lower under reduced temperature and CO2. Seed yield,
fruitcase weight (kernel plus surrounding glume and rachid), and
biomass in 2009e11were higher in the MCC than LGC by an average
of 85%, 99%, and 618%, respectively (Table 4; Fig. 10). In early Holocene conditions, seed yield, weight, and biomass increased by
180%, 206%, and 249% respectively, over those in late-glacial conditions. Fruitcase size varied little between the chambers in all
years and conditions, ranging between about 6.5 and 7.5 mm in
length as in natural teosinte populations today.

domestication process. Two increasingly important areas of


research closely linked to these themes, developmental plasticity
and ecological developmental biology (eco-devo) (West-Eberhard,
2003; Gilbert and Epel, 2009), are key components of the New
Biology and Extended Modern Evolutionary Synthesis, representing the integration of different disciplines and sub-elds not
previously a part of the Modern Synthesis and now thought
necessary to understand the generation of variation and evolutionary change (e.g., Members of the National Research Council,
2009; Pigliucci and Muller, 2010; Wake, 2010). Extending these
concepts to domestication research allows anthropologists to
become fully engaged in what almost certainly will become
important components of the New Modern Evolutionary
Synthesis.
We demonstrated here major phenotypic changes in one generation produced solely through the manipulation of environmental conditions. It is reasonable to expect that the variation we
observed existed in the past climate eras simulated in this experiment; the same branching, inorescence, and stature phenotypic
responses to limiting growing conditions we observed in our study
are seen today in ssp. parviglumis growing in sub-optimal habitats
(Doebley et al., 1995; Whipple et al., 2011), suggesting they are
generalized responses to a variety of limiting growing conditions.
We also note that a ower feminization response to low temperature has been recorded in previous experimental work with Zea and

Table 4
Seed yield and other seed data.

2009
2010
2011
Total
Mean 09e11
2012

Total# viable seeds

Mean# viable seeds per plant

Total weight g

LGC or EHC

MCC

LGC or EHC

MCC

LGC or EHC

MCC

Total# seeds not pollinated or developed


LGC or EHC

MCC

1260
2470
1479
5209
1736 372
5631

2223
2642
4766
9631
3210 787

140
225
123

247
240
397

163 31.4
433

295 51.2
e

47.74
113.11
76.13
237
79 19
242.25

107.88
136.76
225.50
470
157 35
e

230
626
946
1802
601 207
3929

277
986
5085
6348
2116 1498
e

Data are  SEM; For LGC vs. MCC seed number and seed weight 2009e2011, P < 0.001.
Statistical signicance tested by the ManneWhitney rank sum test. Data were not compiled for the MCC in 2012 because branches commonly did not develop owers (see
text).
The LGC is the chamber adjusted to late-glacial conditions in years 2009e11 and the EHC is the chamber adjusted to early Holocene conditions in 2012.

We did not measure photosynthesis or transpiration rate. With


relation to overall seed viability, we recorded the number of fertile
and un-pollinated/undeveloped seeds for each plant (Table 4). The
latter are characterized by being completely white in color in
contrast to the black or other pigmentations on seeds that have
been pollinated and are fertile. There was no evidence for a higher
proportion of defective seeds in the LGC or EHC than MCC. We did
not attempt to formally calculate pollen productivity. However, in
addition to a reduction in the overall number of tassels, tassels had
noticeably fewer main tassel branches in the LGC and EHC, all of
which suggest that pollen grain number was substantially lower
than in the MCC and probably than in todays natural
environments.
5. Discussion
5.1. Phenotypic plasticity and variability
Our research is one of the rst attempts to examine the roles of
plasticity, variability, and the external environment in the plant

a variety of other plant taxa (Richey and Sprague, 1932; HeslopHarrison, 1957), further indicating the importance of temperature
to oral sexuality. The observation in our 2012 study that plants
grown in the modern control chamber from seeds of induced,
maize-like phenotypes reverted back to tall plants with many
branches/branch nodes and branches tipped predominantly by
tassels, and that inbred teosinte seeds behave differently in the two
environmentsdbecoming maize-like phenotypes in the early Holocene conditions e is further evidence of a plasticity response to
the environmental differences.
The data indicate that ssp. parviglumis phenotypes rst exploited and then initially cultivated by human populations differed
substantially from modern plants with a considerable number
possessing maize-type attributes in important vegetative architectural and inorescence sexuality traits, and in seed maturation,
the latter also apparently inuenced by environmental factors. The
implications of our data for understanding teosinte exploitation
and domestication are varied, especially as the kind of detailed
phenotypic information retrieved from this study will not be easily
recovered from early records from maizes homeland due to the

Please cite this article in press as: Piperno, D.R., et al., Teosinte before domestication: Experimental study of growth and phenotypic variability in
Late Pleistocene and early Holocene environments, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2013.12.049

D.R. Piperno et al. / Quaternary International xxx (2014) 1e13

Fig. 7. Right, a mixed maleefemale inorescence with a row of kernels positioned


proximal to the lateral branch (bottom arrow) and male owers terminating the
structure (top arrow). The seeds are not enveloped by vegetative bracts as occurs in
normal teosinte female ears on secondary lateral branches (see Fig. 3). Left, a close-up
of the kernels.

poor preservation of macro-plant remains (seeds, stems, owers)


(Piperno et al., 2009; Piperno, 2011) and scarcity of sites occupied
during the relevant time periods (Ranere et al., 2009; Kennett,
2012).

Fig. 8. Plant height. Notes: *P < 0.001 for LGC vs. MCC mean 2009e2011; P 0.495
(not signicant) for EHC 2012 vs. LGC 09e11; P < 0.001 for EHC 2012 vs. MCC 09e11;
P 0.004 for EHC 2012 vs. MCC 2012. Statistical signicance tested by the Manne
Whitney rank sum test. The error bars represent SD for each individual year and the
SEM for all years. The LGC is the chamber adjusted to late-glacial conditions in years
2009e11 and the EHC is the chamber adjusted to early Holocene conditions in 2012.

For example, plants in our study with maize-like branching,


inorescence, and seed maturation traits were easier and more
efcient to harvest because they have more compact clusters of
female ears located in an easily visible position on the main stem,
and most seeds could be collected with a single harvest effort and
minimal seed loss. Harvesting effectiveness is a central trait inuencing cereal collection and cultivation strategies, a point underscored by the fact that traits associated with harvesting, such as
reduced stem/branch number and uniform seed maturation, are
components of the domestication syndromedthe group of
phenotypic characteristics that were key to initial wild to domesticated transitions (e.g., Olsen and Wendel, 2013). Moreover, the
branch transformations resulted in an increase of a an important
feature called apical dominance that led to a greater concentration
of nutrients in the single main stem, to where female ears were
translocated, eventually permitting the development of large seeds
and cobs of maize (Doebley et al., 1997). Because the maize-like
plants had visible and highly desirable traits with obvious advantages, it is reasonable to expect they would have gained the
attention of early collectors and then cultivators, and the latter
likely would have sought to increase their numbers through
replanting. Their pre-existing availability may have hastened the
successful establishment of Zea cultivation once initiated. However,
indications of short-branched Zea with maize-like oral sexuality
in archaeobotanical records would no longer automatically point to
domestication of these traits, particularly during early cultivation
phases.
Other interesting examples of phenotypic variability occurring
in this study and potentially relevant to the past are the mixed
maleefemale inorescences, many of which had fruitcases lacking
the tightly surrounding vegetative bracts found in normal female
ears (Fig. 7). Exposed fruitcases such as these may also have been
considerably easier to harvest and turn into foods. A possible signicance of these owers in the maize domestication process has
not to our knowledge been explored and may deserve
consideration.

Fig. 9. A box plot of relative branch length vs. inorescence sexuality of primary lateral
branches for data from 2012. For EHC, P < 0.05 for mixed vs. female, P > 0.05 for mixed
vs. male (not signicant), P < 0.05 for male vs. female. Statistical signicance analyzed
by the KruskaleWallis one way analysis of variance on ranks and the Dunns method
for all pairwise multiple comparison procedures.

Please cite this article in press as: Piperno, D.R., et al., Teosinte before domestication: Experimental study of growth and phenotypic variability in
Late Pleistocene and early Holocene environments, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2013.12.049

10

D.R. Piperno et al. / Quaternary International xxx (2014) 1e13

Fig. 10. Plant biomass. Notes: P < 0.001 for LGC vs MCC total mean biomass for 2009e
2011. Statistical signicance tested by the ManneWhitney rank sum test. Error bars
represent the SD.

5.2. The possible role of plasticity in maize domestication


To move forward with a more complete understanding of these
questions requires a better understanding of the links between
genotype, phenotype and the environment. Among the important
avenues of research will be investigating the genetic underpinnings, inheritance, and molecular mechanisms of plasticity;
i.e., if there is underlying genetic variation in the phenotypic responses that we observed to the experimental past environments
and how it responds to selection, together with whether gene
expression changes were involved in the plasticity we recorded. As
discussed above, teosinte plasticity in branching and oral sexuality today, and the differences in these traits between teosinte and
maize, are known to be in part mediated by expression of the tb1
gene, a transcriptional factor that represses bud outgrowth and
increases apical dominance resulting in a maize-like phenotype
(Doebley et al., 1995; Hubbard et al., 2002). Increase in tb1
expression accounts for about 35e50% of the lateral branch number/length and inorescence sexuality changes accounting for the
teosinte to maize transition. It was recently determined that an
allele of tb1 with a transposon insertion called Hopscotch that
enhances the gene expression for short-branch development was
selected during maize domestication (Studer et al., 2011). Hopscotch
is not xed (present) in all modern maize landraces, suggesting
other factors also control plant architecture. It is present as pre-

existing genetic variation in some ssp. parviglumis populations,


suggesting it was available to the rst teosinte cultivators (Studer
et al., 2011). The relationship between Hopscotch and phenotypic
plasticity described here is unknown and should be investigated.
Therefore, questions for future work are many and include
whether an environmental induction (upregulation) of tb1-associated gene expression contributed to the changes we observed and
what other factors might be involved. Whole transcriptome
expression studies (RNAseq) on the plants examined here are in
progress. They will provide a greater understanding of how the
environmentally-induced phenotypic responses we observed
might be coupled with genetic change, and establish an excellent
foundation to explore the connections between ancestral plasticity
and domestication.
In summary, it appears that teosinte foragers and early cultivators worked with wild Zea phenotypes considerably different
than the modern ssp. parviglumis presently used as the baseline for
the domestication process in genetic and archaeological research.
Pre-cultivation availability of maize-like plants may have increased
the speed of the selection process leading to the xation of the
traits, no matter what genetic process resulted in the constitutive
expression of the phenotypes in all environments. Considering that
i) the rst few thousand years of the Holocene in maizes homeland
were not climatically stable, but rather had abrupt 200e400 yearlong reversals in annual precipitation and temperature (c. 2  C
lower) (Bernal et al., 2011), and ii) that until the Industrial Revolution, Holocene atmospheric CO2 was more than 100 ppmv lower
than today, it would be wise to consider when in the Holocene
teosinte became the tall, long-branched plant we observe today.
Finally, although our experiment has opened a wider window
onto the range of phenotypic variation that teosinte collectors and
early cultivators probably saw, there remains much to be learned
from experimental and archaeological study. For example, the
articial selection process that led from an ear of teosinte to a maize
cob is still little understood, and we dont have a good idea of what
early domesticated maize ears were like. The earliest known cobs
dating to 6.7e6.2 ka from Mexico and Peru that survived for
archaeological retrieval only because they came from sites located
in arid environments outside of the origin area and ecological
contexts of maize domestication, are already genetically and
phenotypically well-advanced (Piperno and Flannery, 2001;
Grobman et al., 2012). A major consequence of a long domestication process indicated for maize is that considerable genetic and
possibly phenotypic variation was probably lost. Crop progenitors
faced new environments when taken from their native habitats to
plots prepared for them, and although the latter are characterized
as largely benign compared with natural environments, this does
not mean plasticity responses and the expression of new variation
would have ceased (Schlichting, 2008). This could be especially true
in out-crossing plants with high cryptic or standing genetic variability, such as ssp. parviglumis (Lauter and Doebley, 2002).
5.3. Plant productivity
Previous experimental research indicated signicant decreases
in biomass and seed yield in a range of C3 and C4 wild progenitors of
crops at the Last Glacial Maximum (20 ka) CO2 level of 180 ppmv,
one of the lowest in the history of land plant evolution (Leakey and
Lau, 2012), compared with their growth in Holocene pre-industrial
CO2 levels of 270e280 ppmv (Cunniff et al., 2008, 2010). Seed yield,
which was not studied in teosinte, increased in other C4 species and
by even greater amounts in the C3 plants at 280 ppmv. Photosynthesis itself responded little in teosinte to increased CO2. However,
increasing CO2 to 270 ppmv signicantly lowered the transpiration
rate in teosinte and other C4 species, suggesting that glacial period

Please cite this article in press as: Piperno, D.R., et al., Teosinte before domestication: Experimental study of growth and phenotypic variability in
Late Pleistocene and early Holocene environments, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2013.12.049

D.R. Piperno et al. / Quaternary International xxx (2014) 1e13

water limitation was exacerbated by low CO2 and an improvement


in plant water balance was an important factor that would have
improved teosinte and other C4 plant growth and productivity
during the early Holocene (Cunniff et al., 2008, 2010).
Teosinte grown in this study at late-glacial levels of 206e
215 ppmv CO2 and reduced temperature also responded with
signicantly poorer productivity compared with early Holocene
and modern levels, concordant with previous results. The biomass
and seed yield reductions measured here are more pronounced
than in the Cuniff et al. research for teosinte and other C4 wild
progenitors. The results provide additional evidence that Pleistocene environments imposed physiological stress factors affecting
growth of C4 as well as C3 plants, and indicate that low temperature
was another inhibitory factor on plant growth. Clearly, an important consideration in assessing plant response is the inclusion of
temperature reduction. It is known that different abiotic factors
have interacting effects on plant growth that may not be predictable when one is studied alone, thus lowering CO2 and temperature
simultaneously probably provides a more realistic test of glacial
environments (Cowling and Sage, 1998; Shaw et al., 2002; Ward
et al., 2008; Leakey and Lau, 2012). In a previous experiment on
C3 (Abutilon) and C4 (Amaranthus) plant responses to Pleistocenelike temperature and CO2 combined, the expected, large growth
advantage in biomass measured in the C4 over the C3 plant when
grown in low CO2 alone was considerably attenuated when temperature was simultaneously lowered (Ward et al., 2008). This
nding further indicates that low Pleistocene temperatures had
signicant negative effects on growth of C4 annuals such as
teosinte.
It is important to point out a possible bias in our and other experiments when modern plants are grown for a single generationdthat they do not take into account evolutionary responses
to the past conditions being tested. Multi-year articial selection
studies that measure adaptive responses to Pleistocene factors are
rare but instructive. Ward et al. (2000) found that a C3 annual,
Arabidopsis, partially ameliorated the negative effects of low
(200 ppmv) CO2 on its biomass and seed yield by lengthening the
vegetative growth phase. Single generation experiments may then
overestimate negative responses. However, it is reasonable to
conclude from low CO2 and/or temperature experiments carried
out to date, together with simulation and modeling research on
ecosystem vegetation response to Pleistocene atmospheres and
climate (e.g., Cowling, 2011) and studies on Late Pleistocene fossil
plants that had undergone evolutionary responses to glacial conditions (Ward et al., 2005; Gerhart et al., 2011), that Pleistocene
CO2, temperature, and probably precipitation were signicant
limiting factors on plant growth compared with the following
Holocene. Because plants were adequately watered in this study,
teosinte growth improvement in early Holocene and modern
conditions may have been underestimated.
5.4. Plant productivity, climate change, and agricultural origins
The role of climate change in agricultural origins has long been a
contentious issue in anthropology. The synchronous beginning of
this most fundamental economic transition during the rst few
thousand years of the Holocene in Mesoamerica, South America,
the Near East, and China (e.g. Price and Bar-Yosef, 2011) leads some
scholars, including ourselves, to conclude that the Pleistocenee
Holocene climate and ecological transition resulted in common
underlying inuences on peoples decision to become farmers (e.g.,
Sage, 1995; Richerson et al., 2001; Piperno, 2006, 2011; Cunniff
et al., 2008, 2010). Others have viewed this as an overly deterministic process inadequately focused on socio-cultural factors (see
Piperno and Pearsall, 1998; Price and Bar-Yosef, 2011 for reviews

11

and examples). However, when considered in a broader evolutionary and ecological context, the physical environment becomes
not a simplistic, single-factor or prime mover explanation, but
instead a necessary component of cardinal questions about interactions between humans, their environments, and resource sets
at the transition from foraging to farming.
Our data support hypotheses and previous results bearing on
them which suggest that a persistent cultivation of plants in the
pre-Holocene era may have been difcult to sustain in the face of
low yields of the wild progenitors in Pleistocene environments
(Sage, 1995; Richerson et al., 2001; Cunniff et al., 2008, 2010).
Neither archaeological nor paleoecological data provide estimates
of the relative productivities of crop plant progenitors in preHolocene and Holocene environments, making experimental
research particularly important for assessing the issue. The
archaeological record does bear direct witness to the onset of
cultivation and domestication, and notably, although empirical
data are both rapidly accumulating and being rened around the
world, they continue to indicate that plant food production was
initiated, or at least sustained to the point when it becomes
recognizable and subsequently results in domestication, when the
Holocene began (e.g., Piperno, 2011; Zhao, 2011; Asouti and Fuller,
2012; Willcox, 2013). Experimental productivity and archaeological
data can therefore be used jointly to test hypotheses that limitations on plant growth during the Pleistocene contributed signicantly to the chronology of the rise of agriculture, and as research
progresses hypotheses evaluations will become more robust.
Plant productivity issues are of importance for other hypotheses
of agricultural origins. For example, under the assumptions of
optimal foraging e specically, the diet breadth model e from the
eld of human behavioral ecology (e.g., Kennett and Winterhalder,
2006; Piperno, 2006; Gremillion and Piperno, 2009), increases in
teosinte seed yield and, probably, population density during the
early Holocene would have increased its encounter rate, handling
(collecting) time, resource ranking, and overall foraging efciency,
likely making it more attractive and ultimately adaptively advantageous to human populations who were evaluating and choosing
from the new assortments of resources available to them. This is
one pathway by which previously un- or little-utilized resources
become objects of human attention. Furthermore, it is likely that
potential plant foods differed in their productivity responses
regionally and around the world at the beginning of the Holocene,
and this contributed to the intensication of use of certain species
at the expense of the many others that did not become components
of food producing strategies, as well as to regional chronological
trends and differences in agricultural development (early vs. middle Holocene food production origins). Yield increases would have
also enhanced opportunities for food surpluses and storage, which
in turn may have led to human settlement stability and population
growth (Cunniff et al., 2010).
Reaching a better understanding of how end-Pleistocene environmental shifts affected resource quality together with the
choices people made when they became farmers will involve
extending experimental studies to more plant species. Study has so
far been limited to cereal wild progenitors, but there are numerous
other important ancestral species still common on landscapes
today and available for study. In the New World they include ve
different wild squash (Cucurbita) species and wild legumes such as
Phaseolus common and lima beans, all of which are annuals that
were grown from and for their seeds and should provide rich opportunities for experimental research. Ancient DNA research has
exposed the sometimes incomplete and biased views of domestication history that result from relying on the genetics of modern
domesticated species and their wild ancestors (e.g., Larson et al.,
2007; Roullier et al., 2013). It is clear that we need to also better

Please cite this article in press as: Piperno, D.R., et al., Teosinte before domestication: Experimental study of growth and phenotypic variability in
Late Pleistocene and early Holocene environments, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2013.12.049

12

D.R. Piperno et al. / Quaternary International xxx (2014) 1e13

understand the interplay between past ecology, climate, human


environmental modication, and plant development and phenotypic responses, together with their complex feedbacks. In all elds
concerned with understanding past biota, reconstructions of
vegetation and plant responses to past environmental change have
signicantly depended on modern-day species characteristics and
ecological processes, and rarely provide information on whole
plant attributes of individual species in past era. Thus, experimental
work of this kind may assume increasing importance. Because the
possible range of future plant phenotypic responses may be
signicantly mediated through gene expression and plasticity,
studies such as these also may be important for assessing the effects of global environmental change.

Acknowledgments
Supported by the Smithsonian National Museum of Natural
History, Smithsonian Tropical Research Institute, and a Smithsonian
Institution Scholarly Studies Grant. Special thanks to Eldredge
Bermingham and Cristian Samper for their strong support of the
project. Many thanks to Jorge Aranda and Milton Garca for their
capable help with the plants and environmental chambers, and
Graciela Quijano and Maggie Meagher for their assistance in the
laboratory.

References
Ahn, J., Wahlen, M., Deck, B.L., Brook, E.J., Mayewski, E.A., Andrew, P., Taylor, K.C.,
White, J.W.C., 2004. A record of atmospheric CO2 during the last 40,000 years
from the Siple Dome, Antarctica ice core. Journal of Geophysical Research 109
(D13305), 8.
Asouti, E., Fuller, D.Q., 2012. From foraging to farming in the southern Levant: the
development of Epipalaeolithic and Pre-Pottery Neolithic plant management
strategies. Vegetation History and Archaeobotany 21, 149e162.
Beldade, P., Mateus, A.R.A., Keller, R., 2011. Evolution and molecular mechanisms of
adaptive developmental plasticity. Molecular Ecology 20, 1347e1363.
Bernal, J.P., Lachniet, M., McCulloch, M., Mortimer, G., Morales, P., Cienfuegos, E.A.,
2011. Speleothem record of Holocene climate variability from southwestern
Mexico. Quaternary Research 75, 104e113.
Bush, M.B., Correa-Metrio, A.Y., Hodell, D.A., Brenner, M., Anselmetti, F.S.,
Ariztegui, D., Mueller, A.D., Curtis, J.H., Grzesik, D.A., Burton, C., Gilli, A., 2009.
Re-evaluation of climate change in lowland Central America during the Last
Glacial maximum using new sediment cores from Lake Petn Itz, Guatemala.
In: Vimeaux, F., Sylvestre, F., Khodri, M. (Eds.), Past Climate Variability in South
America and Surrounding Regions. Springer Science & Business Media B.V,
pp. 113e128.
Clark, R.M., Linton, E., Messing, J., Doebley, J.F., 2004. Pattern of diversity in the
genomic region near the maize domestication gene tb1. Proceedings of the
National Academy of Sciences USA 101, 700e707.
Correa-Metrio, A., Bush, M.B., Cabrera, K.R., Sully, S., Brenner, M., Hodell, D.A.,
Escobar, J., Guilderson, T., 2012. Rapid climate change and no-analog vegetation
in lowland Central America during the last 86,000 years. Quaternary Science
Reviews 38, 63e75.
Cowling, S.A., 2011. Ecophysiological response of lowland tropical plants to Pleistocene climate. In: Bush, M.B., Flenley, J.R., Gosling, W.D. (Eds.), Tropical Rain
Forest Responses to Climate Change, second ed. Springer, New York, pp. 359e
380.
Cowling, S.A., Sage, R.F., 1998. Interactive effects of low atmospheric CO2 and
elevated temperature on growth, photosynthesis, and respiration in Phaseolus
vulgaris. Plant Cell and Environment 21, 427e435.
Cunniff, J., Charles, M., Jones, G., Osborne, C.P., 2010. Was low atmospheric CO2 a
limiting factor in the origin of agriculture? Environmental Archaeology 15, 113e
123.
Cunniff, J., Osborne, C.P., Ripley, B.S., Charles, M., Jones, G., 2008. Response of wild C4
crop progenitors to subambient CO2 highlights a possible role in the origin of
agriculture. Global Change Biology 14, 576e587.
Dippery, J., Tissue, D.T., Thomas, R.B., Strain, B.R., 1995. Effects of low and elevated
CO2 on C-3 and C-4 annuals: I. Growth and biomass allocation. Oecologia 101,
13e20.
Doebley, J., 2004. The genetics of maize evolution. Annual Review of Genetics 38,
37e59.
Doebley, J., Stec, A., Gustus, C., 1995. Teosinte branched1 and the origin of maize:
evidence for epistasis and the evolution of dominance. Genetics 141, 333e346.
Doebley, J., Stec, A., Hubbard, L., 1997. The evolution of apical dominance in maize.
Nature 386, 485e488.

Doebley, J., Stec, A., Wendel, J., Edwards, M., 1990. Genetic and morphological
analysis of a maize-teosinte F2 population: implications for the origin of maize.
Proceedings of the National Academy of Sciences USA 87, 9888e9892.
Doust, A.N., 2007. Grass architecture: genetic and environmental control of
branching. Current Opinion in Plant Biology 10, 21e25.
Fusco, G., Minelli, A., 2010. Phenotypic plasticity in development and evolution:
facts and concepts. Philosophical Transactions of the Royal Society B 365, 547e
556.
Gerhart, L.M., Ward, J.K., 2010. Plant responses to low CO2 of the past. New Phytologist 188, 674e695.
Gerhart, I.M., Harris, J.M., Nippert, J.B., Sandquist, D.R., Ward, J.K., 2011. Glacial trees
from the La Brea tar pits show physiological constraints of low CO2. New
Phytolologist 194, 63e69.
Gilbert, S.F., Epel, D., 2009. Ecological Developmental Biology: Integrating Epigenetics, Medicine and Evolution. Sinauer Associates, Inc., Sunderland, MA.
Gremillion, K.J., Piperno, D.R., 2009. Human behavioral ecology, phenotypic
(developmental) plasticity, and agricultural origins: insights from the emerging
evolutionary synthesis. In: Cohen, M.N. (Ed.), Rethinking the Origins of Agriculture, Current Anthropology, vol. 50, pp. 615e619.
Grobman, A., Bonavia, D., Dillehay, T.D., Piperno, D.R., Iriarte, J., Holst, I., 2012.
Preceramic maize from Paredones and Huaca Prieta, Peru. Proceedings of the
National Academy of Sciences USA 109, 1755e1759.
Heslop-Harrison, J., 1957. The experimental modication of sex expression in
owering plants. Biological Reviews of the Cambridge Philosophical Society 32,
38e90.
Hodell, D.A., Anselmetti, F.S., Ariztegui, D., Brenner, M., Curtis, J.H., Gilli, A.,
Grzesik, D.A., Guilderson, T.J., Muller, A.D., Bush, M.B., Correa-Metrio, Y.A.,
Escobar, J., Kutterolf, S., 2008. An 85-ka record of climate change in lowland
Central America. Quaternary Science Reviews 27, 1152e1165.
Hubbard, L., McSteen, P., Doebley, J., Hake, S., 2002. Expression patterns and mutant
phenotype of teosinte branched1 correlate with growth suppression in maize
and teosinte. Genetics 162, 1927e1935.
Iltis, H.H., 1983. From teosinte to maize: the catastrophic sexual transmutation.
Science 222, 886e894.
Iltis, H.H., 1987. Maize evolution and agricultural origins. In: Hilu, K.W.,
Campbell, C.S., Barkworth, M.E. (Eds.), Grass Systematics and Evolution.
Smithsonian Institution Press, Washington, DC, pp. 195e213.
Kennett, D.J., 2012. Archaic-period foragers and farmers in Mesoamerica. In:
Nichols, D.L., Pool, C.A. (Eds.), The Oxford Handbook of Mesoamerican
Archaeology. Oxford University Press, New York, pp. 141e150.
Kennett, D., Winterhalder, B. (Eds.), 2006. Foraging Theory and the Transition to
Agriculture. University of California Press, Berkeley.
Larson, G., et al., 2007. Ancient DNA, pig domestication, and the spread of the
Neolithic into Europe. Proceedings of the National Academy of Sciences USA
104, 15276e15281.
Lauter, N., Doebley, J., 2002. Genetic variation for phenotypically invariant traits
detected in teosinte: implications for the evolution of novel forms. Genetics
160, 333e342.
Leakey, A.D.B., Lau, J.A., 2012. Evolutionary context for understanding and manipulating plant responses to past, present and future atmospheric [CO2]. Philosophical Transactions of the Royal Society B 367, 613e629.
Matsuoka, Y., Vigouroux, Y., Goodman, M.M., Sanchez, J., Buckler, E., Doebley, J.,
2002. A single domestication for maize shown by multilocus microsatellite
genotyping. Proceedings of the National Academy of Sciences USA 99, 6080e
6084.
Members of the National Research Council, 2009. A New Biology for the 21st
Century. Committee on a New Biology for the 21st Century: Ensuring the United
States Leads the Coming Biology Revolution. National Academies Press, Washington, D.C.
Moczek, A.P., 2007. Developmental capacitance, genetic accommodation, and
adaptive evolution. Evolution and Development 9, 299e305.
Moczek, A.P., Sultan, S., Foster, S., Ledon-rettig, C., Dworkin, I., Nijhout, F.,
Abouheif, E., Pfennig, D.W., 2011. The role of developmental plasticity in
evolutionary innovation. Proceedings of the Royal Society B 278, 2705e2713.
Nesbitt, T.C., Tanksley, S.D., 2002. Comparative sequencing in the genus Lycopersicon. Implications for the evolution of fruit size in the domestication of cultivated tomatoes. Genetics 162, 365e379.
Nicotra, A.B., Atkin, O.K., Bonser, S.P., Davidson, A.M., Finnegan, E.J., Mathesius, U.,
Poot, P., Purugganan, M.D., Richards, C.L., Valladares, F., van Kleunen, M., 2010.
Plant phenotypic plasticity in a changing climate. Trends in Plant Science 12,
684e692.
Olsen, K.L., Wendel, J.F., 2013. A bountiful harvest: genomic insights into crop
domestication phenotypes. Annual Reviews of Plant Biology 64, 47e70.
Palmer, C.M., Bush, S.M., Maloof, J.N., 2012. Phenotypic and developmental plasticity
in plants. In: Encyclopedia of Life Sciences. John Wiley&Sons, Ltd, Chichester,
UK.
Phennig, D.W., Mund, M.A., Snell-Rood, E.C., Cruickshank, T., Schlichting, C.D.,
Moczek, A.P., 2010. Phenotypic plasticitys impacts on diversication and
speciation. Trends in Ecology and Evolution 25, 459e467.
Pigliucci, M., Muller, G.B. (Eds.), 2010. Evolution: the Extended Synthesis. MIT Press,
Cambridge, MA.
Piperno, D.R., 2006. The origins of plant cultivation and domestication in the
neotropics: a behavioral ecological perspective. In: Kennett, D., Winterhalder, B.
(Eds.), Foraging Theory and the Transition to Agriculture. University of California Press, Berkeley, pp. 137e166.

Please cite this article in press as: Piperno, D.R., et al., Teosinte before domestication: Experimental study of growth and phenotypic variability in
Late Pleistocene and early Holocene environments, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2013.12.049

D.R. Piperno et al. / Quaternary International xxx (2014) 1e13


Piperno, D.R., 2011. The origins of plant cultivation and domestication in the new
world tropics: patterns, process, and new developments. Current Anthropology
52 (S4), S453eS470.
Piperno, D.R., Pearsall, D.M., 1998. The Origins of Agriculture in the Lowland Neotropics. Academic Press, San Diego.
Piperno, D.R., Flannery, K.V., 2001. The earliest archaeological Maize (Zea mays L.)
from Highland Mexico: new AMS dates and their implications. Proceedings of
the National Academy of Sciences USA 98, 2101e2103.
Piperno, D.R., Moreno, J.E., Holst, I., Lachniet, M., Jones, J.G., Ranere, A.J., Castanzo, R.,
2007. Late Pleistocene and Holocene environmental history of the Iguala Valley,
central Balsas watershed of Mexico. Proceedings of the National Academy of
Sciences USA 104, 11874e11881.
Piperno, D.R., Ranere, A.J., Holst, I., Iriarte, J., Dickau, R., 2009. Starch grain and
Phytolith evidence for early Ninth Millennium B.P. Maize from the Central
Balsas river Valley, Mexico. Proceedings of the National Academy of Sciences
USA 106, 5019e5024.
Price, D., Bar-Yosef, O. (Eds.), 2011. The Origins of Agriculture: New Data, New Ideas.
Current Anthropology, vol. 52. Supplement 4.
Ranere, A.J., Piperno, D.R., Holst, I., Dickau, R., Iriarte, J., 2009. Preceramic human
occupation of the Central Balsas Valley, Mexico; cultural context of early
domesticated Maize and squash. Proceedings of the National Academy of Sciences USA 106, 5014e5018.
Richerson, P.J., Boyd, R., Bettinger, R.L., 2001. Was agriculture impossible during the
Pleistocene but mandatory during the Holocene? A climate change hypothesis.
American Antiquity 66, 387e411.
Richey, F.D., Sprague, G.F., 1932. Some factors affecting the reversal of sex expression
in the tassels of maize. American Naturalist 66, 433e443.
Roullier, C., Benoit, L., McKey, D.B., Lebota, V., 2013. Historical collections reveal
patterns of diffusion of sweet potato in Oceania obscured by modern plant
movements and recombination. Proceedings of the National Academy of Sciences USA 110, 22e25-2210.
Sage, R.F., 1995. Was low atmospheric CO2 during the Pleistocene a limiting factor
for the origin of agriculture? Global Change Biology 1, 93e106.
Schlichting, C.D., 2008. Hidden reaction norms, cryptic genetic variation, and
evolvability. Annals of the New York Academy of Sciences 1133, 187e203.
Shaw, M.R., Zavaleta, E.S., Chiariello, N.R., Cleland, E.E., Mooney, H., Field, C.B., 2002.
Grassland responses to global environmental changes suppressed by elevated
CO2. Science 298, 1987e1990.
Studer, A.J., Doebley, J.F., 2011. Do large effect QTL fractionate? A case study at the
maize domestication QTL teosinte branched 1. Genetics 188, 673e681.

13

Studer, A., Zhao, Q., Ross-Ibarra, J., Doebley, J., 2011. Identication of a functional
transposon insertion in the maize domestication gene tb1. Nature Genetics 43,
1160e1163.
Sultan, S.E., 2010. Plant developmental responses to the environment: eco-devo
insights. Current Opinion in Plant Biology 13, 96101.
Swanson-Wagner, R., Briskine, R., Schaefer, R., Hufford, M.B., Ross-Ibarra, J.,
Myers, C.L., Tifn, P., Springer, N.M., 2012. Reshaping of the maize transcriptome
by domestication. Proceedings of the National Academy of Sciences USA 109,
11878e11883.
Van Heerwaarden, J., Doebley, J., Briggs, W., Glaubitz, J.C., Goodman, M.M., Sanchez
Gonzalez, J., Ross-Ibarra, J., 2011. Genetic signals of origin, spread, and introgression in a large sample of maize landraces. Proceedings of the National
Academy of Sciences USA 108, 1088e1092.
Wake, M., 2010. Development in the real world. Review of ecological developmental
biology: integrating epigenetics, medicine, and evolution by Gilbert, S.F., Epel,
D. American Scientist 98, 75e78.
Wang, H., Nussbaum-Wagler, T., Li, B., Zhao, Q., Vigouroux, Y., Faller, M.,
Bomblies, K., Lukens, L., Doebley, J., 2005. The origin of the naked grains of
maize. Nature 436, 714e719.
Ward, J.K., Antonovics, J., Thomas, R.B., 2000. Is atmospheric CO2 a selective agent
on model C3 annuals? Oecologia 123, 330e341.
Ward, J.K., Myers, D.A., Thomas, R.B., 2008. Physiological and growth responses of
C3 and C4 plants to reduced temperatures when grown at low CO2 of the last ice
age. Journal of Integrative Plant Biology 50, 1388e1395.
Ward, J.K., Harris, J.M., Cerling, T.E., Wiedenhoeft, A., Lott, M.J., Dearing, M.-D.,
Coltrain, J.B., Ehleringer, J.R., 2005. Carbon starvation in glacial trees recovered
from the La Brea tar pits, southern California. Proceedings of the National
Academy of Sciences USA 102, 690e694.
West-Eberhard, M.J., 2003. Developmental Plasticity and Evolution. Oxford University Press, Oxford.
Whipple, C.J., Kebrom, T.H., Weber, A.L., Yang, F., Hall, D., Meeley, R., Schmidt, R.,
Doebley, J., Brutnell, T.P., Jackson, D.P., 2011. grassy tillers 1 promotes apical
dominance in maize and responds to shade signals in the grasses. Proceedings
of the National Academy of Sciences USA 108, 506e512.
Wilkes, H.G., 1977. Hybridization of maize and teosinte in Mexico and Guatemala
and the improvement of maize. Economic Botany 31, 254e293.
Willcox, G., 2013. The roots of cultivation in southwestern Asia. Science 341, 39e40.
Zhao, Z., 2011. New archaeobotanic data for the study of the origins of agriculture in
China. Current Anthropology 52 (S4), S295eS306.

Please cite this article in press as: Piperno, D.R., et al., Teosinte before domestication: Experimental study of growth and phenotypic variability in
Late Pleistocene and early Holocene environments, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2013.12.049

Quaternary International xxx (2014) 1e16

Contents lists available at ScienceDirect

Quaternary International
journal homepage: www.elsevier.com/locate/quaint

Palaeoenvironmental scenarios and lithic technology of the rst


human occupations in the Argentine Dry Puna
Rodolphe Hoguin*, Brenda Oxman
Instituto de Arqueologa, Facultad de Filosoa y Letras, 25 de Mayo 217, 3 Piso, 1002 Buenos Aires, Capital Federal, Argentina

a r t i c l e i n f o

a b s t r a c t

Article history:
Available online xxx

The aim of this research is to contribute to the discussion of environmental scenarios and evaluate in this
context lithic technical strategies developed by hunteregatherer groups during the process of settlement of
the area. The Andean paleoenvironmental knowledge supports the view that during the early Holocene
(10,500e8000 14C BP, uncal.) the environmental conditions were more humid than at present, which would
have produced both an extension of wetlands and an expansion of Andean grassland. However, the results
of pollen analysis in this locality show that these changes were not synchronous. Certain localities may have
retained humid conditions ca. 7000 14C B.P according to the Pastos Chicos record and 7600 14C B.P in the
Lapao record. Thus, the reduction of the distance between the productive patches would have favored a
strategy of highly mobile small groups of hunteregatherers, allowing the supply of raw materials from long
distances, and favoring individual learning, a exible operational chain, and low technical investment. The
Early Holocene is very heterogeneous with numerous environmental and technological changes.
2014 Elsevier Ltd and INQUA.

Keywords:
Paleoenvironmental proxies
Pollen analysis
Lithic technology
Puna Argentina

1. Introduction
Due to the recent development of paleoenvironmental investigations in the area, the tendency has been to make broad
generalizations from a small number of study cases (Oxman and
Yacobaccio, 2014). Nevertheless, recent studies in the Andean
zone indicate environmental variability resulting from the different
response of particular localities to climate changes on a wider scale
(Tchilinguirian and Morales, 2013). Owing to this, it is necessary to
advance the palaeoenvironmental study of the area and evaluate
with precision this differential impact on ancient populations.
This investigation seeks to advance the palaeoenvironmental
studies already performed in the locality (Morales, 2011;
Tchilinguirian et al., 2014a; Oxman and Yacobaccio, 2014;
Tchilinguirian et al., 2014b; among others) by means of new pollen analyses in the Quebrada de Lapao and Pastos Chicos localities
in the Department of Susques, Province of Jujuy, Argentina. In this
way, it will be possible to put in their proper context the rst evidence of human occupations in the area, and to generate a concept
concerning the subsistence strategies of hunteregatherer groups.

* Corresponding author.
E-mail addresses: holocenomedio17cnaa@gmail.com,
(R. Hoguin).

roddh2002@yahoo.fr

More specically, the question of the technical behavior of these


populations during the rst settlement period will be approached.
Secondly, the concept will be contrasted with the results obtained from lithic technology in two sites of the Susques area:
Hornillos 2 and Lapao 9. More specically, this work proposes to
reconstruct operative chains in order to be able to infer technical
and economic behaviors regarding the raw material supply and
lithic technology in a rst settlement context.
1.1. Background research
Traditionally, the advance of the rst human settlement towards
the late Pleistocene, on a macro-regional scale, is associated with
megafaunal hunting (Lynch, 1983; Dillehay et al. 1992; Yacobaccio,
2010, among others). However, in the Central-Southern Andes, no
associations exist between megafauna and materials of anthropic
origin. Chronologically coinciding evidences of megafauna and
human occupation in the area have been discovered, but not at the
same site (e.g. Hippidion sp. at Barro Negro and the archaeological
site Inca Cueva IV) (Yacobaccio and Morales, 2011). No evidences
of Fish-tail type points (Oxman and Yacobaccio, 2014), usually
associated with the hunting of megafauna, have been found, except
one surface example discovered at the Salar Punta Negra 1 site
(Grosjean et al. 2005).
Apparently, in these latitudes, the higher areas (more than 3800
m asl) were colonized later in time, with the retreat of the glaciers.

http://dx.doi.org/10.1016/j.quaint.2014.04.010
1040-6182/ 2014 Elsevier Ltd and INQUA.

Please cite this article in press as: Hoguin, R., Oxman, B., Palaeoenvironmental scenarios and lithic technology of the rst human occupations in
the Argentine Dry Puna, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.010

R. Hoguin, B. Oxman / Quaternary International xxx (2014) 1e16

It is only from ca. 10,500 14C BP when the Puna began to be used
with greater frequency by hunteregatherer groups (Yacobaccio and
Morales, 2011). It has been proposed that the discovered occupations may be the result of a process of dispersion and subsequent
colonization of the area (Yacobaccio, 2010). Then, towards 9500
14
C BP, the evidence suggests that density-dependent mechanisms
will have come into play: demarcation of the space, recurrent circuits, standardization in the use of lithic resources, which would
have increased notably at 8500 14C BP (Yacobaccio and Morales,
2011). In general terms, most early Holocene archaeological sites
(Table 1) are located in caves and rockshelters in gorges and valleys
close to permanent water sources. These settlements would have
low demographic density and high mobility, as shown by the
presence of goods from other ecological zones on both sides of the
cordillera (Nez and Santoro, 1988; Aschero, 1994), and the circulation of obsidian (Yacobaccio et al. 2008). With regards to the
archaeofaunal record, whereas in some sites camelids are an ample
majority, elsewhere it is chinchillids that dominate, the result of
opportunistic hunting on locally available prey (Oxman and
Yacobaccio, 2014).

Table 1
Early Holocene sites in the South-Central Andes.
Site
Tuyajto-1
Tulan-67
Tuyajto-1
Hornillos 2 layer 4
Cueva Yavi layer E, F
Cueva Yavi layer G
Alero Cuevas layer F4
Tambillo 1
Aguas Calientes I-1
Alero Cuevas layer F4
Tambillo-2
Hornillos 2 layer ensemble
Early Holocene
Pintoscayoc 1 Upper layer 6
Salar Punta Negra-1
Pintoscayoc 1 Upper layer 6
Salar Punta Negra-1
Inca Cueva 4
Salar Punta Negra-1
Cueva Yavi
Tambillo-2
Hornillos 2 layer ensemble
Early Holocene
Alero Cuevas layer F4
Inca Cueva 4
Hornillos 2 layer ensemble
Early Holocene
Cueva Yavi
Cueva Yavi layer C
Tuina-5
Inca Cueva 4
San Lorenzo-1
Tuina-5
Huachichocana III
San Lorenzo-1
Pintoscayoc 1 Lower layer 6
Salar Punta Negra-1
San Lorenzo-1
Salar Punta Negra-1
Cueva Yavi layer B
Salar Punta Negra-1
Salar Punta Negra-1
Leon Huasi
Tuln 109
Inca Cueva 4
Pintoscayoc 1 Lower layer 6
Tuina-1

Radiocarbon date
(yrs non cal. BP)

Reference

8130
8190
8210
8280
8320
8420
8504
8590
8720
8838
8870
9150














110
120
110
100
260
70
52
130
100
52
70
50

Nez et al. 2005


Nez et al. 2005
Nez et al. 2005
Yacobaccio et al. 2013
Kulemeyer et al. 1999
Kulemeyer et al. 1999
Lpez 2008
Nez et al. 2005
Nez et al. 2005
Lpez 2008
Nez et al. 2005
Yacobaccio et al. 2013

9180
9180
9190
9230
9230
9450
9480
9590
9590











230
50
110
50
70
50
220
110
50

Hernndez Llosas 2005


Grosjean et al. 2005
Hernndez Llosas 2005
Grosjean et al. 2005
Aschero 2010
Grosjean et al. 2005
Krapovickas 1987e88
Nez et al. 2005
Yacobaccio et al. 2013

9650  100
9650  110
9710  270
9760
9790
9840
9900
9960
10060
10200
10280
10340
10350
10400
10440
10450
10460
10470
10550
10590
10620
10720
10820






















160
100
110
200
125
70
420
120
70
60
130
50
55
50
50
300
150
140
150
630

Lpez 2008
Aschero 2010
Yacobaccio et al. 2013
Krapovickas 1987e88
Kulemeyer et al. 1999
Nez et al. 2005
Aschero 2010
Nez et al. 2005
Nez et al. 2005
Fernndez Distel 1986
Nez et al. 2005
Hernndez Llosas 2005
Grosjean et al. 2005
Nez et al. 2005
Grosjean et al. 2005
Kulemeyer et al. 1999
Grosjean et al. 2005
Grosjean et al. 2005
Fernndez Distel 1989
Nez et al. 2005
Aschero 2010
Hernndez Llosas 2005
Nez et al. 2005

This paper thus begins from the supposition that the effects of
climatic and environmental changes establish the structure of
subsistence resources, which are the bases of decision-making in
hunteregatherers groups (Kelly, 1992). The way in which the
resource structure varies can be understood in terms of predictability, distribution, periodicity, productivity, and the mobility of
the resources, among other factors. In this way it is assumed that
environmental conditions can inuence demographic processes,
which can also restrict or encourage mobility (Binford, 2001). In
turn, changes in mobility and demography can have repercussions
on the networks of information transmission. These networks can
be altered in low demography and population dispersal contexts,
increasing the difculty in the transmission of more complex
techniques (Henrich, 2004). The size of the group composing the
transmission network is important in establishing innovations,
and also the population dispersion and association (Henrich,
2004; Richerson et al. 2009). On the other hand, social structures encompass individuals technical innovations (Roux, 2007).
In this way, it is to be expected that individual learning will predominate when populations are in a dispersed context in new
spaces, not fully resident, as the rst settlers in the Puna (see
Dillehay, 2000; Meltzer, 2003; Yacobaccio and Morales, 2011). In a
process of population dispersal, new habitats and raw materials
are factors which the groups must adapt to. In this context, the
techniques must be simple and exible enough to adapt to
different situations. These technical behaviors are to be expected
in foraging groups with high residential mobility and exploratory
conducts.
Ultimately, so as to deal with lithic evidence in this work, we
might distinguish two kinds of structure: additional structures and
integrated structures (Boda, 2013). The additional structures
possess elements that are independent, in opposition to integrated
structures in which the elements function synergically (Boda,
2013). Additional structures, due to the independence of the
diverse elements, both productive as well as functional aspects of
the techniques, are exible. Thus, this kind of system is to be expected in a context of initial occupation by small groups with high
residential mobility. Therefore, as the ancient populations must
have had to face the ignorance about properties of rocks (block size,
fracture quality, abundance, and so on) found in the region, the
partial transmission of operative chains must have been preferred
as a simple solution.
2. Study area
The study area corresponds to the region of the Dry Puna of
Argentina, between 22 and 24 S and between 3000 and 4500
m asl (Fig. 1). The locality of Susques is located among several
mountains NEeSW oriented mountain chains. The Puna is dened
as a high desert biome, characterized by high solar radiation due to
its high altitude, wide daily thermal amplitude, marked seasonality
in rainfall, and low atmospheric pressure. The vegetation is xerophytic and distributed along an altitudinal gradient, with two main
oristic compositions: tolar vegetation (shrub steppe) and grassland (herbaceous steppe), as well as vegas (wetlands) whose
distribution is azonal (Cabrera, 1976). Several basins with permanent freshwater streams, salt marshes, pans, and beaches (barreales) constitute the drainage system. There are few streams and
watercourses annually available, a critical resource for human
populations in the semi-arid zone (Yacobaccio et al. 2008). The
rainfall (200 mm/year in the region of Susques) occurs mainly in
summer, representing 80% of annual precipitation (Vuille et al.
1998). Altogether, these conditions determine a heterogeneous
distribution of plant and animal resources. Some patches dened as
nutrient concentration zones contain most of the regional biomass

Please cite this article in press as: Hoguin, R., Oxman, B., Palaeoenvironmental scenarios and lithic technology of the rst human occupations in
the Argentine Dry Puna, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.010

R. Hoguin, B. Oxman / Quaternary International xxx (2014) 1e16

available (Yacobaccio, 1994). The most important animal food


sources for humans in the Puna include several mammals (e.g., the
vicua Vicugna vicua and guanaco Lama guanicoe), rodents (eg
vizcachas and chinchillas Chinchilla L. viscacia and brevicaudata),
and cervid (taruca, Hippocamelus antisensis).
At the regional level, it has been argued that environmental
evolution consisted in the passage from more humid and cold nal
Pleistocene conditions to the arid and warm Holocene conditions,
an equivalent to the Younger Dryas (Morales 2010: 102). In the early
Holocene (11,000e8000 14C BP), moisture conditions prevailed in
the high Puna with rainfall 400 mm/year, double the current
200 mm/year (Grosjean et al., 1997). In the Dry Puna, Argentina, the
pollen results presented by Markgraf (1985) from El Aguilar and
Quebrada de Humahuaca suggest a cold and wet environment
between 10000 and 7500 B.P 14C BP, dominated by Poaceae and
herbaceous plants. Although the pre 11,000 B.P 14C and early Holocene records are similar, in Barro Negro receding herbaceous
steppe begins around 11,000 B.P 14C, and in Aguilar (200 m from
Barro Negro) remained until 7500 B.P 14C. The middle Holocene
was characterized by a decrease in rainfall. Several lakes between
20 and 23 S including Titicaca, Salar de Uyuni and the Salar de
Atacama (Bradbury et al., 2001) decreased in level. About 8500
14
C BP, the climate became dry, and between 6000 and 5000 B.P
extreme arid conditions developed. Tchilinguirian (2009) in the
sequence of Laguna Colorada, Catamarca, Argentina, found low
levels of the lake between 7900 and 6300 14C B.P and between 5800
and 4500 14C B.P. However, these conditions were not homogeneous and moist events have been detected in the area. For
example, in the La Hoyada, Catamarca, peat deposits were dated at
8830, 8410 and 8230 14C B.P. (Ratto et al., 2008).

3. Materials and methodology


3.1. Pollen analyses
The pollen analyses consisted of the study of the samples obtained from the proles Lapao 5 and Pastos Chicos, which have
already been analyzed for diatoms, geomorphology and sediment
(Morales, 2011; Tchilinguirian et al., 2014a; Oxman and Yacobaccio,
2014; Tchilinguirian et al., 2014b). From the sedimentary record
of Pastos Chicos (23 400 2900 S; 66 250 3200 W; 3781 m asl), 6 m in
height, 4 samples were taken for dating the prole and 28 samples
for pollen analysis. The bulk organic matter of two of these samples,
PCH2-M2 and PCH1-M3, were dated to 7900  100 14C BP and
8900  130 14C BP using conventional 14C dating (Table 2). Another
date was obtained from a bird bone included in sample PCH2-M15,
dated to 6935  69 14C BP (Fig. 2). Finally in the top of the prole,
PCH2-M17 was dated to 4203  58 14C BP. According to the
radiocarbon dates and the age-depth model (Bennett, 1994) carried
out, they indicate that the prole ranges within a chronology from
ca. 9300 to post-4200 14C BP (Table 3).

Table 2
Sample origin, date and laboratory code.
Sample origin

14C date (Yrs. B.P)

Lab code

Method

Lapao 5 D4
Lapao 5 D3
Lapao 5 D2
Lapao 5 D1
PCH4 M2
PCH1 M3
PCH2 M2
PCH2 M15

7770
8380
8560
9380
4203
8900
7900
6935

LP 981
LP 1518
LP 1763
n/a
AA79835
LP 1841
LP 1836
AA94570

Regular
Regular
Regular
Regular
Regular
Regular
Regular
Regular










80
100
90
110
58
130
100
69

Table 3
Radiocarbon dating and estimated ages, from age-depth model for each of the
samples of Pastos Chicos.
Sample

Depth

PCH2
PCH2
PCH2
PCH2
PCH2
PCH2
PCH2
PCH2
PCH2
PCH2
PCH2
PCH2
PCH2
PCH2
PCH2
PCH2
PCH2
PCH2
PCH2
PCH2
PCH1
PCH1
PCH1
PCH1
PCH1
PCH1
PCH1
PCH1

402
362
331
311
308
280
270
260
250
240
230
220
210
200
190
180
170
168
142
118
118
108
91
63
33
31
20
0

M20
M19
M18
M17 Bis
M17
M16
M15
M14
M13
M12
M11
M10
M9
M8
M7
M6
M5
M4
M3
M2
M8
M7
M6
M5
M4
M3
M2
M1

Chronology

Post 4200

6319
6935  69
6998
7062
7125
7189
7252
7316
7379
7443
7506
7570
7583
7748
7900  100
7900
8015
8210
8532
8877
8900  130
9130
9256

In the case of prole Lapao 5 (23 220 0100 S, 66 210 52,800 W;
3650 m asl), 3 m thigh, 4 samples were taken, three of them 14C
dated (L5_M22 9380  100 14C BP, L5_M15 8560  90 14C BP,
L5_M13 8380  100 14C BP and L5_M8 7770  80 14C BP) and one by
14
C AMS (L5), to date the archive which, according to the age-depth
model (Bennett, 1994) used covers a chronology from ca. 9400 to
7600 14C BP (Fig. 3). In this case, a total of 22 samples have been
processed for pollen analyses (Table 4).

Table 4
Radiocarbon dating and estimated ages, from age-depth model for each of the
samples of Lapao 5. Dates in gray correspond to radiocarbon dating.
Sample

Depth

Chronology

L5M1
L5M2
L5M3
L5M4
L5M5
L5M6
L5M7
L5M8
L5M9
L5M10
L5M11
L5M12
L5M13
L5M14
L5M15
L5M16
L5M17
L5M18
L5M19
L5M20
L5M21
L5M22
L5M23

15
30
80
85
120
140
150
185
210
220
240
260
270
300
310
325
330
345
350
360
365
370
380

7375
7447
7483
7519
7591
7627
7734
7770
7878
7949
8164
8236
8380
8500
8560
8660
8800
8840
8920
9060
9080
9380
9340

 80

 100
 90

110

Please cite this article in press as: Hoguin, R., Oxman, B., Palaeoenvironmental scenarios and lithic technology of the rst human occupations in
the Argentine Dry Puna, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.010

R. Hoguin, B. Oxman / Quaternary International xxx (2014) 1e16

Fig. 1. Map of localization of early Holocene sites and study area.

The methodology in the pollen analyses adhered to the standard


protocol for Quaternary pollen (Faegri and Iversen, 1989). The
laboratory stage consisted in the observation of the samples under
a Zeiss-Axiolab biological microscope, counting at least 200 grains
per sample whenever possible.
Identication of the pollen types has been performed using the
bibliography and available atlases for the area under study
(Heusser, 1971; Markgraf and DAntoni 1978) and the reference
catalog of the Palynology laboratory of the Faculty of Agricultural
Sciences UNJu/Conicet. For the statistical analysis of the data, the
TILIA program (Grimm, 1987, 2004), was used, with an analysis of
main components (CONISS) which allowed the discrimination of
two principal palynological zones.
The palynological interpretation is based on ecological criteria
of actualism and uniformitarianism. Therefore, in this case it is
possible to use the information of current vegetation for making
palaeoenvironmental interpretations. In this sense, descriptions of
vegetation belts and ecologically related matters in the Puna, carried out by different scholars, among them Cabrera (1976), Ruthsatz
and Movia, 1975 were used.
On the other hand, several paleoecological studies from the
Altiplano have suggested that increases in Poaceae and decreases in
Asteraceae indicate wetter conditions (and vice versa) (Liu et al.
2005). The Poaceae/Asteraceae proxy follows a precipitationdriven vegetation gradient on the Altiplano, in which grasses
dominate in the northern (wetter) sections of the region, while
Asteraceae shrubs dominate in the drier southern sections. This
ecological relationship forms the basis for our interpretation of the
pollen record. Therefore, we can use the logarithmic P/A ratio as a

humidity index for the Altiplano. Accordingly, the P/A ratio would
be 0 if the Poaceae and Asteraceae pollen percentages are equal.
Positive numbers show the dominance of grasses and therefore
wetter conditions. Negative values suggest the dominance of the
Asteraceae over grasses, and therefore drier conditions (Liu et al.
2005).
3.2. Lithic technology
Operative chains are reconstructed in the framework of lithic
technology. One operative chain is dened as the logical and
organized concatenation of technical steps, from the supply of raw
materials to the discarding of artifacts, including all stages of production and use of the tools (Inizan et al. 1995). This methodology
allows choices, possibilities, and concepts of volumetric reduction
to be put in evidence (Boda, 2013). In this way, after determining
operative schedules, hypotheses can be drawn regarding the conducts, knowledge, know-how, and skills of the stone-knappers
(Inizan et al. 1995), and consequently the mechanisms of cultural
transmission (Roux, 2007).
In lithic technology, at least two aspects can be considered with
regard to the level of integration (to determine whether a structure
is additional or integrated): production and function (or rather
techno-function). In production, the level of integration depends on
the relation between the volume of the exploited raw material and
the remaining (unused) volume. The larger the exploited volume in
relation to the remaining one, the greater is the level of integration
in production (Boda, 2013). Concerning the techno-functional
aspect, a tool will be considered integrated in which both techno-

Please cite this article in press as: Hoguin, R., Oxman, B., Palaeoenvironmental scenarios and lithic technology of the rst human occupations in
the Argentine Dry Puna, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.010

R. Hoguin, B. Oxman / Quaternary International xxx (2014) 1e16

Fig. 2. Sediment reconstruction of Pastos Chicos prole. Pastos Chicos was the sequence studied by pollen analysis.

functional units (prehensile, transformative, and energytransmitting TFUs) and their shaping (Boda, 2013). It is important to evaluate the level of integration and technical investment in
the production of blanks (i.e. knapping), in the shaping of the tools,
and the degree of integration between them.
The rst step in any lithic analysis, particularly the technological, consists in analyzing the separate assemblages raw materials.
Three categories will be taken into account: 1) the akes from the
debitage; 2) the debris of the shaping; and 3) the tools themselves.
From this classication, it will be possible to have an idea of the
activities carried out at the sites, as well as strategies of supply and
transport of raw materials, which can also hint at the mobility of
the groups.
Subsequently, knapping procedures will be analyzed by producing diacritical diagrams of the cores, showing order and direction in the extractions, and of the blanks, as much in their
extraction as in their dimensions. Diacritical diagrams of the tools
will be made in order to determine their shaping procedures,
showing the series that allowed the Techno-functional Units to be
established. The interdependence among the different stages of the
operative chains, determined by the carrying out of preliminary
stages (preparation) and by the identication of the different series
of blanks and akes from cores and tools reduction, will allow the
knapping-procedures to be characterized as additional or
integrated.
The sample analyzed comes from the assemblage of the initial
early Holocene layers at Hornillos 2 (layers 6, 6A, B, C, and D), from

layers 5 and 4 of the same site (nal early Holocene), and from a
small above-surface assemblage at the Lapao 9 prole, possibly
corresponding to nal early Holocene occupation (Table 5). Layer 5,
a clay silt sandy lens layer, has an intermediate situation in the
stratigraphy, and is without dating (Yacobaccio et al., 2013). It
contains some similar artifacts to those of layer 4. For this reason, it
probably corresponds to a nal early Holocene occupation. In this
work, only tools from this level were analyzed.

Table 5
Sample per site.

Tools
Non retouched
artifacts
Cores

Layers 6, 6A, B, C & D


Hornillos 2

Layer 5
Hornillos 2

Layer 4
Hornillos 2

Lapao 9

31
3771

3
46

32
2488

3
0

4. Results
4.1. Pollen analyses
In the case of Pastos Chicos, 17 samples were subjected to pollen
analysis. Fifteen taxa have been identied, grouped into large

Please cite this article in press as: Hoguin, R., Oxman, B., Palaeoenvironmental scenarios and lithic technology of the rst human occupations in
the Argentine Dry Puna, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.010

R. Hoguin, B. Oxman / Quaternary International xxx (2014) 1e16

Fig. 3. Sediment reconstruction of Lapao 5 prole. Lapao5 was the sequence studied by pollen analysis.

categories that agree with the present composition of Puna vegetation: the herbaceous steppe represented by Poaceae and Ephedrae; and the shrub steppe represented by the Asteraceae,
Fabaceae, Rosaceae, Solanaceae and Mimosaceae taxa.
Chenopodiaceae-Amaranthaceae, and Urticaceae were identied as
disturbance indicators. Finally, the indicators of local humidity
were also determined, including Cyperaceae, Pteridophytes, Halorgaceae and Carex sp. The fern spores, fungal spores, and algae
colonies are excluded from the pollen sum.
From the analysis it has been possible to detect two different
intervals on the basis of the composition of the vegetation: 1) between 9300 and 7000 14C BP, when a stable steppe grassland was
recorded (represented mainly by the Poaceae family between 80
and 100%), with pollen elements dened as indicators of local humidity (v.s.); and 2) after 7000 to 4200 14C BP, a clear change in
plant composition is evident, in which a decrease in the herbaceous
steppe is observed (Poaceae family) accompanied by a gradual increase of the shrub steppe (mainly the Asteraceae family). Isolated
humidity events have also been recorded around 6300 14C BP
(Fig. 4). From the humidity index applied, only towards the period
after 6300 14C BP (sample P2 M17) did it show a negative value,
which is interpreted as the most arid interval in the whole
sequence (Fig. 5).
In the case of Lapao 5, 15 samples were subjected to pollen
analysis. A total of 15 taxa have been identied. Representing the
grass steppe are Poaceae and Ephedra sp. Representing shrub
steppe are Asteraceae, Mimosaceae, Halorgaceae, Chuquiraga sp.
and Fabiana sp. The following taxa are identied as local humidity

indicators : Tagetes sp., Myriophyllum quitense, pteridophytes,


Nototriche sp, Cyperaceae, and various types of fungal spores. Pollen
types from trees such as Alnus acuminata (from the yunga) have
also been found. Pollen types have been grouped that, found
together, may be good indicators of the anthropogenic impact, i.a.:
Malvaceae, Urticaceae, Chenopodiaceae and Amaranthaceae. Fern
spores, fungal spores, and algae colonies are excluded from the
pollen sum.
It is possible to register two different periods in taxa composition: 1) between 9280 and 8400 14C BP, in which a mixed steppeland vegetation of grasses and bushes is observed (mostly the
Poaceae and Asteraceae families) with intervals with high percentages of M. quitense, interpreted as an indicator of local humidity and low temperatures, with a maximum between 8600 and
8400 14C BP (M. quitense, Cyperaceae, Nototriche sp., Tagete sp.;
among others). Diatom studies also show a dry phase around 8400
14
C BP detected by an increase of the littoral species (Tchilinguirian
et al., 2014a); 2) From 8400 to 7600 14C BP, a fall is observed in the
local humidity indicators. However, the presence of a water body is
still evident. Conditions of aridity set in towards 7600 14C BP, with
the drying out of the wetland (Fig. 6). From the application of the
humidity index, negative values have been detected around 8840
14
C BP (sample M18) and 9080 14C BP (sample M21), which are
interpreted as moments of particular aridity (Fig. 7). The results
conrm the preliminary pollen results and the palaeoenvironmental interpretations presented for the same localities in
previous works (Morales, 2011; Tchilinguirian et al., 2014a; Oxman
and Yacobaccio, 2014; Tchilinguirian et al., 14b).

Please cite this article in press as: Hoguin, R., Oxman, B., Palaeoenvironmental scenarios and lithic technology of the rst human occupations in
the Argentine Dry Puna, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.010

Fig. 4. Pollen diagram of Pastos chicos prole.

R. Hoguin, B. Oxman / Quaternary International xxx (2014) 1e16

Please cite this article in press as: Hoguin, R., Oxman, B., Palaeoenvironmental scenarios and lithic technology of the rst human occupations in
the Argentine Dry Puna, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.010

R. Hoguin, B. Oxman / Quaternary International xxx (2014) 1e16


Table 6 (continued )
Site

Tambillo 1
Tulan-67
Toconce

Fig. 5. Moisture index of Pastos Chicos prole (Log P/A).

4.2. Lithic technology


In northern Chile and northwest Argentina, triangular stemless
weapon-heads predominate for use in individual hunting strategies
(Pintar, 1995; Aschero and Martnez, 2001; De Souza, 2004; Nez
et al. 2005; among others). In Chile, two phases were distinguished
for the early Holocene: Tuina (11,000e9500/9000 14C BP), characterized by the presence of eponymous weapon-heads and a raisedback scraper; and Tambillo (9500/9000e8500/8000 14C BP), characterized by the presence of cupuliform points (see Nez and
Santoro, 1988). It is not the objective of this work to use these
phases as operative units for this investigation, but to contextualize
the diagnostic artifacts that might possibly be identied in the area
under investigation. An important superposition between the
contexts with a simultaneous presence of these diagnostic artifacts
is noticeable, with an early tendency for Tuina and Tambillo towards the end of the early Holocene (Table 6, Fig. 8).

Table 6
Early Holocene sites, radiocarbon dates and frecuencies of Tuina and Tambillo
points.
Site

Radiocarbon date
Tuina Tambillo References
(yrs BP no calibrated)

Tuln 109

10590  150

Salar Punta Negra-1 10470


10460
10440
10350






50
50
50
60

9450
9230
9180
10820
10060
9840
10400
10280
9960
10620
9900
9650
92330
9710
9590
9150
8280



















50
50
50
630
70
110
130
120
125
140
200
110
70
270
50
50
100

8720
8210
8130
8870






100
110
110
70

Tuina-1
Tuina-5
San Lorenzo-1

Inca Cueva 4

Hornillos 2 layer
ensemble Early
Holocene
Hornillos 2 layer 4
Aguas Calientes I-1
Tuyajto-1
Tambillo-2

Nez
et al. 2005

Grosjean
et al. 2005

2
2

e
e

Nez et al. 2005


Nez et al. 2005

e
3

Nez et al. 2005


8

Hocsman
et al. 2012

13
14

e
4

Yacobaccio
et al., 2013
Hoguin 2013 Ms
Yacobaccio
et al., 2013
Nez et al. 2005
Nez et al. 2005

Nez et al. 2005

7
e

Radiocarbon date
Tuina Tambillo References
(yrs BP no calibrated)
9590
8590
8190
7990






110
130
120
125

e
1
e

8
6
2

Nez et al. 2005


Nez et al. 2005
Nez and
Santoro 1988

At Hornillos 2 (Fig. 9), the relation between tools and debitage


akes is reverted between the sum of layers from the beginning of
the early Holocene (Hornillos 2 layers 6, 6A, B, C, and D) and the
layer from the end of early Holocene (layer 4 at Hornillos 2). In the
rst assemblage, in quartzite (directly local <1 km) and andesite
(intermediately local <30 km), the tools are represented in a
smaller proportion than are the akes, whereas for the obsidian
(non-local >90 km) and various silica (local <10 km) their representation is greater. In layer 4 the opposite phenomenon is found:
quartzite and andesite tools are more highly represented in relation
to the akes of these raw materials, whereas obsidian and the silica
are to a lesser degree. Finally, in layer 4, andesite is better represented than in the diverse layers of the beginning of the early
Holocene at Hornillos 2, and the proportion of obsidian tools is
lower.
From the cores discovered and from the identied blanks (unshaped) it is possible to infer main debitage procedures at the start
of the early Holocene. One consists in the frontal unidirectional
exploitation from a single scar front (Fig. 10:1), and the other in an
alternating centripetal exploitation (Fig. 10: 2). For this period,
akes from unidirectional extractions dominate. Although there are
no cores for layer 4 at Hornillos 2, the analysis of the akes, shaped
as well as unmodied and their comparison with other sites (see
Hoguin, 2013) indicates a variability of blanks, as well as a
complexity of debitage procedures (Fig. 11). It is possible to determine a unidirectional method such as occurred previously, but the
existence of blanks which characteristics (platforms and removals)
indicates that the cores were mostly exploited, with the incorporation of new stages which alternated extraction surfaces and
platforms (Fig. 11). In some cases, a third platform was used to
obtain overspilling akes, with an abrupt back (Fig. 11). In this case,
obtaining this kind of blank would depend on the previous series of
removals. In this way, greater diversication can be observed in the
types of blanks obtained, with the presence of dihedral-but akes
(exploitation alternating surfaces and platforms) and nucleus
anks (Fig. 11).
As at other sites in Chile, Tuina points are preferentially present
in the layers from the beginning of the early Holocene, and the
Tambillo points are preferentially present in the layer from the
end of the early Holocene. Two shaping procedures were identied for Tuina points. The rst consists of bifacial shaping in two
interdependent sequences. The second consists in shaping by hierarchical treatment of the surfaces, with partial alternate nishing retouches (Fig. 12). In this last case the different sequences do
not seem dependent on each other, and may reect different lifestages of the artifact. In the cases in which it was possible to
identify the technical axis, this did not coincide with the
morphological axis. These observations were also made at Inca
Cueva 4 and Alero Cuevas (Hocsman et al., 2012; Hoguin and
Restifo, 2012). This might indicate independence between the
production of blanks and the shaping of the tools, as well as a low
predetermination during the debitage procedure. These observations can be extrapolated to other tools that are not projectilepoints (Fig. 12: 6 and 7). For layer 4 at Hornillos 2 and the Lapao

Please cite this article in press as: Hoguin, R., Oxman, B., Palaeoenvironmental scenarios and lithic technology of the rst human occupations in
the Argentine Dry Puna, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.010

R. Hoguin, B. Oxman / Quaternary International xxx (2014) 1e16

Fig. 6. Pollen diagram of Lapao 5 prole.

9 sequence, we can observe on Tambillo points a shapingprocedure by hierarchical treatment of the surfaces in three or
four sequences, possibly interdependent, as they do not seem to
correspond with reactivation stages, and with coinciding negative

bulbs on the cutting edges (Fig. 13). Layer 4 additionally contains


other types of projectile-points (Fig. 14), among them a few fragments of Huiculunche 2 points identied at several locations in
Chile and Argentina (De Souza, 2004; Nez et al. 2005; Hoguin,
2013 Ms).
During the early Holocene, various tools show 2 transformative
TFUs that were made in 3 unifacial sequences (Fig. 15). The blanks
employed vary in their dimensions, and may have been obtained by

Fig. 7. Moisture index of Lapao 5 prole (Log P/A).

Fig. 8. Seriation of Tuina and Tambillo points (using logarithmic values frequencies
with the Past version 2b17b Spindle Diagram).

Fig. 9. Raw material proportions for tools and debitage akes. A: Early Holocene layers
ensemble of Hornillos 2; B: Hornillos 2 layer 4.

Please cite this article in press as: Hoguin, R., Oxman, B., Palaeoenvironmental scenarios and lithic technology of the rst human occupations in
the Argentine Dry Puna, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.010

10

R. Hoguin, B. Oxman / Quaternary International xxx (2014) 1e16

Fig. 10. Idealized debitage procedures and its products for initial early Holocene.

the unidirectional recurrent method. However, some of these tools


were manufactured with agstones as blanks (Fig. 15: 3). There
appears to be independence between the techno-functional objectives and the production of blanks. In layer 4 at Hornillos 2, towards the end of the early Holocene, various types of tools and
various shaping-procedures were observed. Some show a 4-stage
shaping process of two different sorts of TFUs. One is a sinuous
cutting-edge made by alternating bifacial shaping, and the other
(an abrupt cutting-edge) by unifacial shaping in two sequences
(Fig. 16). Finally, there is another artifact class showing a hierarchic
treatment of the surfaces (Fig. 17).
5. Discussion
During the early Holocene, herbaceous vegetation is recorded
w500 m below its present location, characteristic of higher and
more humid zones. This increase in regional humidity would have
brought about an extension and increase of productive patches, as
well as that of associated critical resources and a reduction of the
distance between them, which may have enlarged the capacity to
support animal biomass. These conditions may be explainable
owing to rises in temperature in relation to the previous period,
accompanied by a rise in rainfall, cloudiness, and therefore a

decrease in evapotranspiration. In turn, the maintenance of these


grasslands can also be attributed to water produced by the thawing
of glaciers (Nez and Grosjean, 1994).
Regional conditions towards greater aridity set in ca. 8000
14
C BP though certain localities may have retained humid conditions to ca. 7000 14C BP according to the Pastos Chicos record. In the
Quebrada de Lapao the retreat of humid conditions began ca. 8400
14
C BP, and the apparent disappearance of the water body ca. 7600
14
C BP. These conditions are compatible with the observations at
Pastos Chicos (Tchiliguirian et al., 2014a). Thus, the presence of
grassland can be detected until ca. 7000 14C BP (Tchilinguirian et al.,
2014b).
The changes to more arid conditions in the mid-Holocene were
not synchronous, which would indicate the existence of productive
patches of different quality and availability during the early Holocene. This would represent a change in the segmentation of the
space, with patches of differential quality and productivity.
The expectations arising from this climatic context, and for this
stage of settlement of the area ca. 10,200 and 9300 14C BP (see
Yacobaccio and Morales, 2011), would be the predominance of individual learning and exible operative chains. This gradual adaptation process, biological as well as cultural, will have been
followed by greater complexity of the operative chains and

Please cite this article in press as: Hoguin, R., Oxman, B., Palaeoenvironmental scenarios and lithic technology of the rst human occupations in
the Argentine Dry Puna, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.010

R. Hoguin, B. Oxman / Quaternary International xxx (2014) 1e16

11

Fig. 11. Idealized debitage procedures and its products for nal early Holocene.

stabilization in the networks for the transmission of technical


knowledge, specially visible from 8500 14C BP, as mechanisms
dependent on density came into action (see Yacobaccio and
Morales, 2011).

For the rst occupations in Susques, the lithic archaeological


evidence that allows the proposed expectations to be contrasted,
corresponds to a lapse between ca. 9700 and 9100 14C BP, which
would reect an averaged archaeological record. This would be the

Fig. 12. Tuina weapon head type.

Please cite this article in press as: Hoguin, R., Oxman, B., Palaeoenvironmental scenarios and lithic technology of the rst human occupations in
the Argentine Dry Puna, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.010

12

R. Hoguin, B. Oxman / Quaternary International xxx (2014) 1e16

Fig. 13. Tambillo weapon head type.

product of diverse events, of dispersal as much as colonization


(Dillehay, 2000; Yacobaccio, 2010) of the Puna. This context is
characterized by the presence of Tuina points, in which the shaping
design can be variable. The blanks for these points seem to have

been the product of an ad hoc selection, as the technical axis never


corresponds with the morphological. This could indicate a low
predetermination in the production of blanks, and independence
between the production and tool project. As shown, other tools also

Fig. 14. Other projectile points types.

Please cite this article in press as: Hoguin, R., Oxman, B., Palaeoenvironmental scenarios and lithic technology of the rst human occupations in
the Argentine Dry Puna, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.010

R. Hoguin, B. Oxman / Quaternary International xxx (2014) 1e16

13

Fig. 15. Initial early Holocene processing tools types.

present the same characteristics. Other larger-sized artifacts have


two perpendicular TFUs shaped in two or three sequences, and
were fashioned from diversied blanks (in some cases made
directly with agstones), possibly also chosen ad hoc. These tools
may correspond to what some authors called raised-backed
(Nez and Santoro, 1988), due to their thickness or in some cases
the presence of a thick projecting ridge, obtained prior to the
shaping of these tools. Several of the blanks for these tools are
obtained from a type C debitage procedure, thus an additional
structure (Boda, 2013). There is a certain independence between
blank production schedules and the objectives of their nal carving,
allowing technical exibility.
In layers 5 and 4 at Hornillos 2, the latter dated at ca. 8300
14
C BP, a greater diversity of artifacts is observed, including projectile points as in other tools. The shaping process of Tambillo
points differs from those of Tuina. Through the shaping protocols
of the different tools from layer 4, clearer patterns have been
detected than in the previous layers (6, 6A, B, C, and D), as well as
greater technical investment. This possibly involved an increase in
the learning time and greater skills, with more systematized
knapping projects, which may have been maintained by biased
transmission in a context of more stable relationships between
individuals.
The diversity observed in layer 4 could be the result of an
increase in technical innovations relative to the previous period.

Although the debitage procedures are also of additional structure, they present a greater degree of integration in the productive stages and higher complexity from the incorporation of
certain predetermined stages and objectives such as obtaining
cores ank akes. From the evidence at other sites, it was
possible to propose that there might be a certain dependence
between the debitage and shaping processes of tools for this
period (Hoguin, 2013).
In the same way, the changes observed between both contexts
on the proportions of the different raw materials and the indices of
the tool-debris relationship, might reect changes in supply strategies from ca. 8500 14C BP. They could be the result of systematization in the supply of raw materials from an intermediate distance,
such as andesite (w20e30 km) with blanks. This could reect
diminished residential mobility of the groups, resulting from more
permanent occupations in higher quality resource patches
(Aschero, 1994; Morales, 2011).
6. Conclusions
Pre-Holocene conditions seem not to have permitted human
occupation, at least on a sustained basis (Yacobaccio and Morales,
2011; Tchilinguirian et al., 2014a). Neither has evidence been
found of associations between human occupations and megafauna,
despite their contemporaneity in the region. It has been proposed

Please cite this article in press as: Hoguin, R., Oxman, B., Palaeoenvironmental scenarios and lithic technology of the rst human occupations in
the Argentine Dry Puna, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.010

14

R. Hoguin, B. Oxman / Quaternary International xxx (2014) 1e16

Fig. 16. Final early Holocene processing tools types.

that the earliest settlement of land above 4000 m asl would have
taken place during a span of ca. 900 years, and towards 9387  18
14
C BP all the Puna living spaces had already been occupied by the
hunteregatherer groups (Yacobaccio and Morales, 2011). These
permanent occupations will have been possible due to the stabilization of the resource patches (Aldenderfer, 1999), as palaeoenvironmental data generated for the early Holocene in the area
seem to show. The archaeofaunal evidence suggests the resources
were highly available locally in a humid environment and were
obtained close to the sites (Yacobaccio and Morales, 2011), which is
associated with opportunist hunting of locally available resources
(Oxman and Yacobaccio, 2014).
In this context, it is possible to conclude that the groups stabilized rapidly, at least from the evidence available at Hornillos 2. In
addition, the choice of pigments for the rock-art at Hornillos 2 and
Inca Cueva 4 reects provisioning from local ranges, without precluding interaction (Yacobaccio et al. 2008). Although at the very
start of the colonization the use of Puna space may have been
exclusively temporary owing to the restrictions of biological

adaptation to the region, as from 9500 14C BP, it is conceivable that


the groups exclusively occupied the Puna (in the case of the study
area), following a North-to-South mobility, as the supply of obsidians and pigments shows (Yacobaccio et al. 2008; Yacobaccio and
Morales, 2011).
In addition, it is possible to show that the technical changes
that took place ca. 8300 14C BP are synchronous with the climate
changes. All these changes could also be related to a longer
residence of population in this environment, a better knowledge
of the landscape, and greater stability of their transmission
networks. The new conditions of aridness could have had consequences on the organization of the populations, on the strategies of resource provisioning in general, and on hunting in
particular, as shown by the gradual rise in the representation of
camelids towards the end of the early Holocene at various sites
(Yacobaccio et al., 2013). These changes may have fostered the
rise of the innovations observed in the archaeological record
through a diversication of artifacts, as observed at the Hornillos
2 rockshelter.

Please cite this article in press as: Hoguin, R., Oxman, B., Palaeoenvironmental scenarios and lithic technology of the rst human occupations in
the Argentine Dry Puna, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.010

R. Hoguin, B. Oxman / Quaternary International xxx (2014) 1e16

15

Fig. 17. Tools with hierarchic treatment faces shaping (Hornillos 2 layer 4).

Acknowledgments
Thanks to Miguel Eduardo Burbano and Javier Francisco Aceituno to their invitation at this issue, to Hugo Yacobaccio, Eric Boda
and Liliana Lupo, to CONICET PIP 3173 project and Universit de
Paris X, CNRS UMR7041. We also want to thank the reviewers of this
work for their comments and suggestions.

References
Aldenderfer, M., 1999. The Pleistocene/Holocene transition in Peru and its effects
upon human use of the landscape. Quaternary International 53 (54), 11e19.
Aschero, C.A., 1994. Reexiones desde el Arcaico Tardo (6.000-3.000 A.P). Rumicatana Ao 1, 13e17.
Aschero, C.A., Martnez, J.G., 2001. Tcnicas de caza en Antofagasta de la Sierra, Puna
Meridional Argentina. Relaciones de la Sociedad Argentina de Antropologa
XXVI, 215e241.
Bennett, K.D., 1994. Condence intervals for age estimates and deposition times in
late-Quaternary sediment sequences. The Holocene 4, 337e348.
Binford, L.R., 2001. Constructing Frames of Reference. An Analytical Method for
Archaeological Theory Building Using Ethnographic and Environmental Data
Sets. University of California Press, London.
Boda, E., 2013. Techno-logique et technologie. Une palo-histoire des objets lithiques tranchants. @rcho-ditions.com.
Bradbury, J.P., Grosjean, M., Stine, S., Sylvestre, F., 2001. Full- and Late-glacial lake
records along PEP-1 transect: their role in developing inter-hemispheric paleoclimate interactions. In: Markgraf, V. (Ed.), Interhemispheric Climate Linkages.
Academic Press, San Diego, pp. 265e291.
Cabrera, A.L., 1976. Regiones Fitogeogrcas Argentinas. Editorial Acme, Buenos
Aires.
De Souza, P., 2004. Cazadores Recolectores del Arcaico Temprano y Medio en la
cuenca superior del ro Loa: Sitios, conjuntos lticos y sistemas de asentamiento.
Estudios Atacameos 27, 7e43.
Dillehay, T., 2000. The Settlement of the Americas: a New Prehistory. Basic Books,
New York.
Dillehay, T.D., Ardilla Calderdn, G., Politis, G., de Moraes Coulinho Beltrao, M. da C.,
1992. Earliest hunters and gatherers of South America. Journal of World Prehistory 6, 145e204.
Faegri, K., Iversen, J., 1989. In: Faegri, K., Kaland, P.E., Krywinski, K. (Eds.), Textbook
of Pollen Analysis. John Wiley and Sons, Chichester.
Fernndez Distel, A., 1986. Las Cuevas de Huachichocana, su posicin dentro del
precermico con agricultura incipiente del Noroeste argentino. In: Beitrage Zur

Allegemeinen und vergleichenden Archaeologie, Band 8. Verlag Phillip von


Zabern, Mainz Am Reim, pp. 353e430.
Fernndez Distel, A., 1989. Una nueva cueva con maz acermico en el N. O.
Argentino: Len Huasi 1, excavacin. Comunicaciones Cientcas 1, 4e17.
Grimm, E., 1987. CONISS: a FORTRAN 77 program for stratigraphically constrained
cluster analysis by the method of incremental sum of squares. Computers and
Geosciences 13 (1), 13e35.
Grimm, E., 2004. Tilia and TGView 2.0.2 Illinois State Museum. Research and
Collection Center, Springeld, Illinois.
Grosjean, M., Nez, L., Cartajena, I., Messerli, B., 1997. Mid-Holocene climate and
culture change in the Atacama desert, northern Chile. Quaternary Research 48,
239e246.
Grosjean, M., Nez, L., Cartajena, I., 2005. Palaeoindian occupation of the Atacama
Desert, northern Chile. Journal of Quaternary Science, 643e653.
Henrich, J., 2004. Demography and Cultural Evolution: why adaptive cultural processes produced maladaptative losses in Tasmania. American Antiquity 69 (2),
197e214.
Hernndez Llosas, M.I., 2005. Pintoscayoc and the archaeology of the arid puna and
rift valley, northern Argentina. In: Smith, H. (Ed.), 23 South, Archaeology and
Environmental History of the Southern Deserts. National Museum of Australia
Press, Canberra, pp. 186e197.
Heusser, C.J., 1971. Pollen and Spores of Chile. Modern Types of the Pteridophyta,
Gymnosperma, and Angiospermae. The University of Arizona Press, Tucson.
Hocsman, S., Martnez, J.G., Aschero, C.A., Calisaya, A.D., 2012. Variability of Triangular Non-Stemmed Projectile Points of Hunter-Gatherers of the Argentinian
Puna. Current Research in the Pleistocene. Southbound: Late Pleistocene
Peopling of Latin America, 63e67.
Hoguin, R., Restifo, F., 2012. Patterns of Cultural Transmission in the Manufacture of
Projectile Points: Implications for the Early Settlement of the Argentine Puna.
Current Research in the Pleistocene. South Bound. Late Pleistocene Peopling of
Latin America, 69e73.
Hoguin, R., 2013. Evolucin y cambios tcnicos en sociedades cazadoras recolectoras
de la Puna Seca de los Andes Centro-Sur. Tesis de doctorado en Arqueologa de
la Facultad de Filosofa y Letras, Universidad de Buenos Aires, Universit de
Paris X.
Inizan, M.-L., Reduron, M., Roche, H., Tixier, J., 1995. Technologie de la pierre taille,
Tome 4. CREP, Meudon.
Kelly, R.L., 1992. Mobility/sedentism: concepts, archaeological measures, and effects. Annual Review of Anthropology 21, 43e66.
Krapovickas, P., 1987e1988. Noticia. Nuevos fechados radiocarbnicos para el sector
oriental de la Puna y la Quebrada de Humahuaca. Runa XXVIIeXVIII, 207e219.
Kulemeyer, J.A., Lupo, L.C., Kulemeyer, J.J., Laguna, L.R., 1999. Desarrollo paleoecolgico durante las ocupaciones humanas del precermico del norte de la
Puna Argentina. In: Schbitz, F., Liebricht, H. (Eds.), Bejtrge zur quartren
Landschaftsentwicklung Sdamerikas. Bamberger Geographische Schriften,
Bamberg, pp. 233e255.

Please cite this article in press as: Hoguin, R., Oxman, B., Palaeoenvironmental scenarios and lithic technology of the rst human occupations in
the Argentine Dry Puna, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.010

16

R. Hoguin, B. Oxman / Quaternary International xxx (2014) 1e16

Liu, K.B., Carl, A.R., Thompson, L.G., 2005. Ice-core pollen record of climatic changes
in the central Andes during the last 400 yr. Quaternary Research 64, 272e278.
Lpez, G.E., 2008. Arqueologa de Cazadores y Pastores en Tierras Altas: Ocupaciones humanas a lo largo del Holoceno en Pastos Grandes, Puna de Salta. BAR
International Series, Argentina. Oxford.
Lynch, T.F., 1983. The Paleo-indians. In: Jennings, J.D. (Ed.), Ancient South Americans. W.H. Freeman and Co, San Francisco, pp. 87e137.
Markgraf, V., 1985. Paleoenvironmental history of the last 10,000 years in northwestern Argentina. Zentralblatt fur Geologie un Palaontologie 11, 1739e1749.
Markgraf, V., DAntoni, H.L., 1978. Pollen ora of Argentina. The University of Arizona Press, Tucson.
Meltzer, D.J., 2003. Lessons in landscape learning. In: Rockman, M., Steele, J. (Eds.),
Colonization of Unfamiliar Landscapes. The Archaeology of Adaptation. Routledge, London, pp. 222e241.
Morales, M.R., 2011. Arqueologa ambiental del Holoceno temprano y medio en la
Puna Seca Argentina. British Archaeological Reports. In: South American
Archaeology Series, vol. 15. Archaeopress, Oxford.
Nez, L., Grosjean, M., 1994. Cambios Ambientales Pleistocnico-Holocnicos:
Ocupacin Humana y Uso de Recursos en la Puna de Atacama (Norte de Chile).
Estudios Atacameos 11, 11e24.
Nez, L., Santoro, C., 1988. Cazadores de la Puna Seca y Salada del rea Centro Sur
Andina (Norte de Chile). Estudios Atacameos 9, 13e65.
Nez, L., Grosjean, M., Cartajena, I., 2005. Ocupaciones Humanas y Paleoambientes
en la Puna de Atacama. Taraxacum: Instituto de Investigaciones Arqueolgicas y
Museo Universidad Catlica del Norte.
Oxman, B.I., Yacobaccio, H.D., 2014. Pollen analysis of Pastos chicos: palaeoenvironmental and archaeological implications during the Holocene in the
Dry Puna of Argentina. In: Kligmann, D.M., Morales, M.R. (Eds.), Physical,
Chemical and Biological Markers in Argentine Archaeology: Theory, Methods
and Applications, British Archaeological Research International Series.
Archaeopress, Oxford (in press).
Pintar, E., 1995. Los conjuntos lticos de los cazadores Holocnicos en la Puna Salada.
Arqueologa 5, 9e24. Instituto de Ciencias Antropolgicas, FFyL, UBA.
Ratto, N., Montero, C., Hongn, F., Randall, M., 2008. Gente y volcanes: el registro
arqueolgico de ambientes inestables del oeste Tiogasteo de Catamarca (ca.
5000e1500 AP). In: Primeras jornadas de arqueologa del rea Punea de los
Andes centro-sur, Horco Molle, Tucumn, Argentina, pp. 113.

Richerson, P.T., Boyd, R., Bettinger, R.L., 2009. Cultural innovations and demographic
changes. Human Biology 81, 211e235.
Roux, V., 2007. Ethnoarchaeology: a non historical science of reference necessary
for interpreting the past. Journal of Archaeological Method and Theory 14, 153e
178.
Ruthsatz, B., Movia, C., 1975. Relevamiento de las estepas andinas del noreste de la
provincia de Jujuy. FECYT, Argentina.
Tchilinguirian, P., 2009. Paleoambientes holocenos en la Puna austral, Provincia de
Catamarca (27 S): Implicancias geoarqueolgicas. Ph.D. thesis. Facultad de
Ciencias Exactas y Naturales, Universidad de Buenos Aires.
Tchilinguirian, P., Morales, M.R., Oxman, B., Lupo, L.C., Olivera, D.E., Yacobaccio, H.D.,
2014a. Early to Middle Holocene transition in the Pastos Chicos record,dry Puna
of Argentina. Quaternary International 330, 171e182.
Tchilinguirian, P., Morales, M.R., 2013. Mid-Holocene paleoenvironments in
Northwestern Argentina: main patterns and discrepancies. Quaternary International 307, 14e23.
Tchilinguirian, P., Morales, M.R., Oxman, B., Pirola, M., 2014b. Paleoenvironmental
studies of the Quebrada de Lapao, Jujuy Province, Argentina (23 220 0100 S, 66
210 52,800 W, 3650 m a.s.l.) for the 9400e7300 yrs B.P. span. In: Kligmann, D.M.,
Morales, M.R. (Eds.), Physical, Chemical and Biological Markers in Argentine
Archaeology: Theory, Methods and Applications, British Archaeological
Research International Series. Archaeopress, Oxford (in press).
Vuille, M., Hardy, D.R., Braun, C., Keimig, F., Bradley, R.S., 1998. Atmospheric circulation anomalies associated with 1996/1997 summer precipitation events on
Sajama ice cap, Bolivia. Journal of Geophysical Research 103, 11191e11204.
Yacobaccio, H.D., 1994. Biomasa animal y consumo en el Pleistoceno-Holoceno
Surandino. Arqueologa 4, 43e71.
Yacobaccio, H.D., 2010. The Paleoindian and Archaic of Central and South America.
In: Renfrew, C., Bahn, P. (Eds.), The Cambridge World Prehistory.
Yacobaccio, H.D., Cat, M.P., Sol, P., Alonso, M.S., 2008. Estudio arqueolgico y
fsico-qumico de pinturas rupestres en Hornillos 2 (Puna de Jujuy). Estudios
Atacameos 36, 5e28.
Yacobaccio, H.D., Morales, M.R., Sol, P., Samec, C.T., Hoguin, R., Oxman, B., 2013.
Mid-Holocene occupation in the Dry Puna in NW Argentina: evidence from the
Hornillos 2 rockshelter. Quaternary International 307, 38e49.
Yacobaccio, H.D., Morales, M., 2011. Ambientes pleistocnicos y ocupacin humana
temprana en la Puna argentina. Boletn de Arqueologa PUCP 15, 337e356.

Please cite this article in press as: Hoguin, R., Oxman, B., Palaeoenvironmental scenarios and lithic technology of the rst human occupations in
the Argentine Dry Puna, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.010

Quaternary International xxx (2014) 1e13

Contents lists available at ScienceDirect

Quaternary International
journal homepage: www.elsevier.com/locate/quaint

Detailed fatty acids analysis on lithic tools, Cerro El Sombrero Cima,


Argentina
Natalia Mazzia*, Nora Flegenheimer
CONICET-rea Arqueologa y Antropologa, Municipalidad de Necochea, Av. 10 y calle 63 s/No , 7630 Necochea, Buenos Aires, Argentina

a r t i c l e i n f o

a b s t r a c t

Article history:
Available online xxx

This paper describes results obtained through fatty acids analysis on a sample of lithic artifacts. Analyzed
tools come mainly from the excavated assemblage of Cerro El Sombrero Cima (Tandilia, Argentina),
where the occupation has been assigned to the Pleistocene/Holocene transition. They include shtail
projectile points (FTPP), retouched tools, and ground tools. A variety of resources have been identied
which indicate the great diversity of organic materials used and highlights the importance of plants in
past daily life. Also, through this method marine resources have been identied, giving support to the
proposition that different environments, including the sea coast, were familiar to early hunter-gatherers
in the region. In addition, results are relevant to the discussion of the hafting and recycling of FTPP and
the generalized use of other tool types.
2014 Elsevier Ltd and INQUA. All rights reserved.

Keywords:
Pampean region
Lithic artifacts
Organic resources
Gas chromatography

1. Introduction

2. Regional environment

Many early archaeological sites in the Argentine Pampas have


poor preservation conditions for organic material. Such is the case
of Cerro El Sombrero Cima (CoSC). This site, with the largest early
lithic assemblage identied, has both poor stratigraphy and low
organic preservation. This situation has promoted the exploration
of an analytical methodology that provides information about the
organic resources on which lithic artifacts were used. Fatty acids
analysis by gas chromatography has proved successful in recovering valuable information in the absence of macroscopic evidence.
It thus constitutes a useful approach to understanding the resources used by people in the past and the way in which tools were
employed.
This paper describes the results obtained on a sample of artifacts, mainly from the excavated assemblage of CoSC. It includes
shtail projectile points (FTPP), retouched tools, and ground tools.
A variety of resources have been identied which are relevant to
discuss the movements of early people and the environments they
visited, as well as the resources which were part of their daily lives.

The case study is located in a hilly area in the center east of


Buenos Aires province, currently approximately 80 km from the
Atlantic coast (Fig. 1). These hills belong to the Tandilia ranges, one
of the two mountain systems that interrupt the extensive Pampean
plains. Tandilia consists of low hills crossing the plains over 350 km
with a northwest-southeast orientation, which form groups
disconnected by uvial erosion (Zrate and Rabassa, 2005). In the
section under study, the igneous and metamorphic Precambriam
bedrock is overlain by late Precambrian sedimentary rocks and
Palaeozoic quartzites (Zrate et al., 1993) forming typical buttes and
mesas with at summits. Water can be found mainly in springs and
small streams which ow from the hills to the ocean. One major
river, the Quequn Grande, collects water from Tandilia, traverses
the southern plains and ows into the ocean. Water is also available
in seasonal lagoons in the plains and in some coastal lagoons. The
plain ends in a wide chain of sand dunes surrounding the seashore
or in coastal cliffs up to 30 m high where the Tandilia ranges meet
the Atlantic coast.
The Pampean landscape is covered by a loess mantle, nowadays
rich agricultural land. The latest cycle of eolian sedimentation
began during the last glaciation. Eolian deposition decreased
dramatically in the area during the early Holocene when more
stable conditions favored soil development. Loess deposits have
been intensely affected by pedogenesis and biological activity, as
well as by aqueous transport (Flegenheimer and Zrate, 1993;
Zrate et al., 1993). The Tandilia ranges exhibit a great diversity of

* Corresponding author.
E-mail addresses: natymazzia@yahoo.com.ar (N. Mazzia), noraf@necocheanet.
com.ar (N. Flegenheimer).
http://dx.doi.org/10.1016/j.quaint.2014.04.027
1040-6182/ 2014 Elsevier Ltd and INQUA. All rights reserved.

Please cite this article in press as: Mazzia, N., Flegenheimer, N., Detailed fatty acids analysis on lithic tools, Cerro El Sombrero Cima, Argentina,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.027

N. Mazzia, N. Flegenheimer / Quaternary International xxx (2014) 1e13

Other environmental issues relevant to this paper are the location of the marine coast and the geomorphology of the Rio de la
Plata during the time of human occupation. At this time, the sea
level would have been lower than today, some 60 m below its
current position (Guilderson et al., 2000; Parker et al., 2008; Ponce
et al., 2011). This modied the coastline extending the plains
mainly in the north and the south of the province of Buenos Aires,
especially at the latitude of the Ro de la Plata. Modications were
less dramatic in the central portion, where our study area is situated, which was then w120 km inland. According to recent studies
(Ponce et al., 2011), the occupation took place between two moments of inferred stabilization (between buried marine Terrace I
and II). For that time, the rate of recession of the coastline has been
estimated at an average of 27 m per year: therefore, changes in the
sea coast must have been visible within a human lifetime. The
different coastline affected the river distribution and drainage
network. In this scenario, the Ro de la Plata discharged eastwards
from its present position, was narrower than today (Cavallotto
et al., 2002), and its current course was mostly a coastal plain.
There were no major geographical barriers between the plains in
Uruguay and Buenos Aires province (Fig. 1).

Fig. 1. Map of the study area.

microenvironments, and early archaeological occupations have


been registered in many of them. Sites at higher elevations are set
in environments with low eolian sedimentation rates, which may
have been stabilized prior to human occupation. Early occupations
located on the hill slope and in more protected rockshelters probably took place during the end of the depositional event and before
the early Holocene interval of soil development. This stable interval
was interrupted by the arid mid Holocene episode (Zrate and
Flegenheimer, 1991; Zrate et al., 2000).
A grass steppe environment has been proposed for the time of
early human occupation in the region (Paez et al., 2003). Recently,
some indicators of the presence of shrub or arboreal components
have been recorded at archaeological sites. The presence of charcoal indicates woody plants, and isolated micro particles related to
Tala trees (Celtis tala) have been found in the hills and the Quequn
Grande valley (Mazzanti, 2003; Martnez and Gutirrez, 2011;
Flegenheimer et al., 2013a).
The stratigraphic resolution in the valleys is greater than in the
interuves, with thick Holocene sequences in the oodplains of
major rivers where alluvial sedimentation rates were higher
(Zrate et al., 2000). Paleoclimatic conditions have been intensely
studied in these sequences. In the Quequn Grande, several climatic
pulses have been recorded and uctuating climatic conditions
inferred. A stable interval, following early human occupations, led
to the development of the regional paleosol Puesto Callejn Viejo
(Zrate et al., 2000; Gutirrez et al., 2011; Martnez et al., 2013)
which has been recently dated at Paso Otero archaeological locality
giving ages of 10,000e9400 BP (Johnson et al., 2012).
Paleontological evidence suggests that since the Middle Pleistocene and through part of the Holocene, cold and arid climatic
conditions prevailed in the region, alternating with more humid
periods (Tonni et al., 1985). By the time of human occupation
addressed in this paper a great variety of extant and extinct species
were included in the assemblages with bone preservation, with
some such as Equus (Amerhippus) neogaeus indicating arid conditions (Mazzanti, 2003; Martnez et al., 2013). The decrease of
megamammals in the following millennia has promoted discussion
about the nature of human intervention in their extinction in the
region, in view of the absence of evidence indicating mass killings
(Martnez et al., 2013).

3. Archaeology, case study


Hunter-gatherer groups inhabited the Pampean Region since
the Late Pleistocene. Based on faunal remains, a generalized
economy has been inferred for these early societies. Both extinct
and extant species show signs of having been consumed and used
(Mazzanti, 2003; Martnez and Gutirrez, 2004; Martnez et al.,
2013). Among them, some such as Lama guanicoe, Ozotoceros
bezoarticus, Hemiauchenia sp., Hippidion sp. and Equus sp. that show
anthropic intervention can reasonably be assumed to have been
hunted with FTPP, although hunting sites with clear associations
are elusive in the region. Information about the use of vegetal resources is scarcer, and except for the presence of charcoal, it is
mainly indirect and comes from use wear studies.
Currently, twenty six archaeological sites assigned to the Late
Pleistocene and Early Holocene bear witness of these early occupations: Cerro La China 1, 2 and 3, Cerro El Sombrero Cima and
Abrigo 1, Los Helechos, Cueva Zoro, El Ajarafe, El Mirador, Abrigo La
Grieta, Arroyo Seco 2, Paso Otero 5, Campo Laborde, La Moderna,
Cueva Tixi, Abrigo Los Pinos, Cueva La Brava, Cueva El Abra, Cueva
Burucuy, Amalia 2, Loberia 1, El Guanaco 1 and 2, Arroyo de Fras,
Laguna El Doce, and Pehuen Co paleoichnological site
(Flegenheimer, 2003; Mazzanti, 2003; Bayn et al., 2004; Politis,
2008; Messineo and Politis, 2009; Mazzanti et al., 2010; Bayn
et al., 2011; Martnez and Gutirrez, 2011; Mazzia, 2011, 2013;
Politis et al., 2011; vila, 2011; Mazzanti et al., 2012; Mazzia and
Flegenheimer, 2012). Dates from these sites range between ca.
12,000 and 9000 14C BP; in the microregion under study, sites
rmly correlated to this case study range between 10,000 and
11,000 14C BP (Mazzia and Flegenheimer, 2012).
Cerro El Sombrero Cima (CoSC) is an open air site located on the
summit of El Sombrero hill, a butte where an early assemblage has
been collected both from the surface and in stratigraphy
(Flegenheimer, 2003; Flegenheimer et al., 2013b). These last remains are included in an A soil horizon which overlies either a B
paleosol horizon or the quartzitic bedrock. No organic remains have
been recovered at the site, as the sedimentary layer has been
subject to active pedological activity and preservation is very poor.
No radiocarbon dates associated to the occupation are available,
which has been assigned to the PleistoceneeHolocene transition by
comparison with nearby dated sites (Zrate et al., 2000/2002).
The assemblage under study includes 1501 aked tools, of
which 90 are shtail projectile points (FTPP) in different moments

Please cite this article in press as: Mazzia, N., Flegenheimer, N., Detailed fatty acids analysis on lithic tools, Cerro El Sombrero Cima, Argentina,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.027

N. Mazzia, N. Flegenheimer / Quaternary International xxx (2014) 1e13

of their use life, and 11 tools manufactured by pecking, and


grinding, one of them decorated by engraving. Toolstone has been
highly selected, the most frequently chosen raw material being
colored orthoquartzite of the Sierras Bayas Group (SBGO) found in
quarries 40e60 km distant (Bayn et al., 1999; Colombo and
Flegenheimer, 2013). Also, quartz, chert, silicied limestone,
locally available orthoquartzites, dacite, silicied dolomite, and
other rocks have been found in small proportions. Most of these
rocks are exposed in the Tandilia ranges. Quarries for chert and
silicied dolomite have been studied (Flegenheimer, 1991; Barros
and Messineo, 2004, 2006) and potential sources for quartz and
local orthoquartzites are known (Colombo, 2013). The only long
distance transport identied involves silicied limestone, which
probably traveled within social interaction networks. Its sources
have been located 400e500 km to the northeast, crossing the Ro
de la Plata (Flegenheimer et al., 2003).
CoSC exhibits several characteristics which make it an exceptional site. The view from the hilltop is panoramic and reaches
40 km in most directions. It is denser and larger than most other
early sites. The tool assemblage includes infrequently found tool
types, the breakage ratio is very high (90%), fractures affect FTPP as
well as other artifact types, bifacial tools are more frequent than in
other regional early sites (42%), and debitage is restricted to the last
moments of tool manufacture, resulting mainly from bifacial
aking with soft hammer percussion (Weitzel, 2012; Flegenheimer
et al., 2013b). Microscopic use wear analysis on a small sample of
tools from the site was useful to identify the importance of wood
used by early peoples in an environment where this resource must
have been scarce (Leipus, 2010).
The site is interpreted as a lookout and a place chosen for
refurbishing weapons and discarding tools broken elsewhere. It has
been recently proposed that this place was meaningful for early
settlers and relevant in non-verbal communication. This proposal
has been further sustained by comparison with another distant site,
Cerro Amigo Oeste, which reveals a similar choice in social landscape (Flegenheimer et al., 2013b).

4. Gas chromatography analysis of lithic artifacts


4.1. Lipids and gas chromatography analysis
Organic residue analysis makes use of analytical organic
chemical techniques to identify the presence and the origins of
organic remains that cannot be described using traditional
methods of archaeological investigation (Evershed, 2008). In this
case study, this kind of analysis allows us to recover valuable information about organic resources which have not left macroscopic
evidence. The study of the absorbed substances may provide information about dietary aspects or use of artifacts in the past
(Evershead et al., 1992; Evershed, 2008). They reveal different kind
of organic resources utilized and, when possible, their provenance.
When processing the archaeological objects, a small sample of
fat and/or oil is obtained, that was trapped in the pores and ne
cracks of the rocks while the objects were being used. Lipids, such
as fats and oils, are organic molecules present in plant and animal
tissues, composed of carbon, hydrogen, and oxygen atoms. These
molecules are hydrophobic. They exhibit stability under high
temperatures and a minimum breakdown over time in constant
environmental conditions (Feiser and Feiser, 1960; Evershed, 1993;
Gunstone et al., 2007). Thus, lipids can survive absorbed in the
matrix of stone artifacts. In addition, due to their hydrophobic
property, there is a minimum loss of substances when objects are
washed with water in the absence of soap or detergents (Evershed,
1993; Babot, 2004).

Fatty acids are one of the natural compounds of lipids, with a


linear chain and an even number of carbon atoms. They are classied according to chain length and the presence/absence of carbon
double bonds. They are often divided into saturated and unsaturated forms: the carbon chains in saturated fatty acids are fully
substituted with hydrogen, while unsaturated fatty acids have one
or more double bonds between carbon atoms (Feiser and Feiser,
1960; Fankhauser, 1994; IUPAC, 1997; Bondia Pons, 2007;
Gunstone et al., 2007).
The deterioration of polyunsaturated and very long chain fatty
acids is more pronounced, but saturated fatty acids are less affected
by decomposition. In archaeological samples, the presence of the
rst ones is important, as they indicate some stability of the substances over time (Malainey et al., 1999).
Gas chromatography is a widely used technique for the separation, identication, and quantication of lipids. It consists in the
separation of mixtures of volatile or semi-volatile organic compounds through the use of protocols and specic equipment.
4.2. Methodological precautions
For this study, several methodological precautions were taken in
order to ensure a more accurate interpretation of the results. First,
as skin lipids may possibly be transferred to the artifacts, affecting
the results of the chromatographic analysis (Evershead et al., 1992),
a control sample was included. The aim was to consider whether a
ake without use exhibited fatty acids due to its exposure to
environmental conditions and our manipulation. According to the
results, the archaeological samples under study are not polluted by
manipulation during excavation or laboratory tasks (Mazzia, 2010e
2011).
Second, since lipids are constituent parts of the environment, it
is possible to mistakenly interpret lipids absorbed from the burial
environment as evidence of past resources processed with
archaeological artifacts (Evershed, 1993; Buonasera, 2007). Bearing
in mind this possible contamination source, we also analyzed
sediments of the A soil horizon where the assemblage was recovered. Based on the differences in the lipid record exhibited by the
artifact and sediment samples, we consider that there is minimal
post-burial absorption of substances. This can be explained in
terms of the hydrophobic nature of the lipids that limits their
migration by dissolution or dissemination. Fatty acids from the soil
may be absorbed in the porosities of rocks, but this process is so
slow that they are affected by degradation prior to accumulating in
appreciable amounts (Charters et al., 1993; Evershed, 1993;
Buonasera, 2007). These two methodological precautions provide
greater certainty to the assessment of past organic substances in
the archaeological samples.
Finally, a small experimental lithic collection was prepared in
order to analyze absorbed substances from different origins.
Experimental SBGO akes were used with meat (Bos taurus), plants
(including a mix of edible grass leaves, bulbs, roots, and fruits),
wood (Salix sp.), vegetal mastic (made with heated Pinus sp. resin
and ashes), and a mixture of meat and plants (Mazzia, 2010e2011).
4.3. Materials and method
The samples were processed at the Chemical Analysis Lab of
Materia Hnos. Oleochemichals (Mar del Plata, Argentina). Artifacts
with sizes between 2 and 10 cm were totally immersed in chloroform in a covered beaker for 24 h with an intermittent shakeup. As
extraction involved the complete artifact, the fatty substances obtained come from the whole stone tool, not only from the edges. As
a consequence, results may include a mix of substances coming
from different edges and surfaces. Larger objects were treated

Please cite this article in press as: Mazzia, N., Flegenheimer, N., Detailed fatty acids analysis on lithic tools, Cerro El Sombrero Cima, Argentina,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.027

N. Mazzia, N. Flegenheimer / Quaternary International xxx (2014) 1e13

differently, as only selected parts were in contact with the solvent.


After this rst step, every extract was ltered and dried and treated
under nitrogen. In all cases, the extraction process took place at a
constant room temperature with nitrogen supply in order to prevent possible deterioration of lipids.
After the extraction of samples, a protocol of methylation was
followed. Methylation involves the preparation of methyl esters
derived from the samples of lipids. The protocol (EC 2-66 preparation of methyl esters) is based on the ofcial methods of the
American Oil Chemists adapted by the Materia Hnos. Lab. It
included the following steps:
 the extract was diluted in 12.5 ml of a methyl solution of potassium hydroxide in a 150 ml ground-necked round-bottom
ask,
 the round-bottom ask was connected to a refrigerant while
boiling for 12 min,
 25 ml of a 5% solution of sulphuric acid in methanol were added,
 the round-bottom ask was connected to a refrigerant while
boiling for 12 min,
 it was cooled,
 for the phase separation, a 500 ml separatory funnel was used
with 100 ml of distilled water and 50 ml of petroleum ether; the
upper phase containing the methyl esters was kept,

 this extract was collected in a 100 ml beaker,


 the solvent was boiled off using a double boiler and completely
dried in a 90  C oven,
 it was diluted with petroleum ether and transferred into a vial,
 2 ml of the sample were injected into the gas chromatograph.
When it is possible, the use of gas chromatographyemass
spectrometry equipment is preferable to gas chromatographs
because it is a more accurate technology for the identication of
pure substances (balos et al., 2003). However, gas chromatographs are widely available and also offer a reliable and effective
method for identication and quantication of fatty acids feasible
for correlation with natural resources (Malainey et al., 1999;
Malainey, 2007; Costa Angrizani and Constenla, 2010). Although
on this occasion the samples were processed with gas chromatographs, previously we had used gas chromatographyemass spectrometry equipment for other samples (Babot et al., 2007; Mazzia,
2010e2011). According to our experience with both types of
equipment, the results have not shown signicant differences.
In this case, a HP 6890N gas chromatograph with ame ionization detector (FID) and a back automatic injector were used for
the analysis, along with a Supelcowax 10 capillary column. The
column temperature was 200  C, the injector temperature was
250  C, and the detector temperature was 280  C.

Fig. 2. Four chromatograms presented as examples: A. sample of sediments; B. S12 304 39 (fragment of an unidentied polished artifact); C. S13 905 3 (FTPP recycled as a drill); D.
S7 102 1 (fragmented side scraper).

Please cite this article in press as: Mazzia, N., Flegenheimer, N., Detailed fatty acids analysis on lithic tools, Cerro El Sombrero Cima, Argentina,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.027

N. Mazzia, N. Flegenheimer / Quaternary International xxx (2014) 1e13

The next step consisted in determining the percentage of all the


fatty acids in the chromatograms resulting from the injections
(Fig. 2). These percentages and the relations between them were
used for interpreting the results. The presence of some fatty acids,
mainly saturated and short-chain ones, may be due to the action of
degradation processes that have led to the disappearance of most
volatile fatty acids (Malainey et al., 1999). These fatty acids are
expected to be abundant in degraded samples.
The identication of substances is based on chemist-taxonomic
principles. This implies relating a chemical property in the sample,
such as the presence or absence of a typical compound or mixture
of compounds, with that same property of contemporary plant and
animal products taken as reference (Evershead et al., 1992). Thus,
when interpreting each sample, the relative percentages of fatty
acids were compared with the composition of animal fats and
vegetable oils that are described in current databases, experimental
information or archaeological papers (for example: U.S. Testing
Company, N/D; Patrick et al., 1985; Rottlnder, 1990; Robinson
et al., 1991; A&G Tcnica, 1993; Fankhauser, 1994; Caabate
Guerrero and Snchez Vizcano, 1995; Fezler, 1995; Brenner and
Bernasconi, 1997; Pond et al., 1997; Malainey et al., 1999; Sengr
et al., 2003; Abd El-Baky et al., 2004; Babot et al., 2007; Buonasera, 2007; Frre et al., 2010; Mazzia, 2010e2011; Muhamad and
Mohamad, 2012).
Among the materials recovered at Cerro El Sombrero Cima,
different types of stone tools were selected for the analysis of the
substances that could have adhered to their surfaces. The analyzed
assemblage includes 29 lithic objects: ve pecked and ground stone
tools, nine fragmented FTPP, and 16 aked artifacts such as scrapers

and bifaces (Figs. 3e5). This selection includes a variety of artifacts


favoring pecked and ground stone tools which were expected to
preserve fatty acids according to previous work on this and other
assemblages (Babot, 2004; Babot et al., 2007; Flegenheimer et al.,
2013a). Also, several FTPP were analyzed, as these are the most
conspicuous artifacts in the assemblage. In this group, points which
were recycled into other tools were favored as they were expected
to yield more information about a variety of resources used. Finally,
in the third group, tools were chosen to represent the morphological variety exhibited at the hilltop. Most tools were recovered
from within the A soil horizon at the site except for a small fragmented sphere (206) which is the only surface remain, and was
included to test possible fatty acid preservation under unfavorable
conditions. Most of the tools come from the centre of the hilltop in
the main excavation area (S12 and S13), although examples from
test pits in other areas of the hilltop were included (W130 and S7).
5. Results
The results are organized into three groups according to the
characteristics of the analyzed objects (Tables 1e3). Each of the
tables includes a rst column containing the results of the gas
chromatographic analysis of sediments: even though this column
appears in the three tables, it describes the same sample. The lipid
record in the sediment sample consists of seven fatty acids and 6.6%
undetermined compounds. In the chromatogram (Fig. 2A), the
peaks that represent these fatty acids have small areas. However, as
mentioned in the following descriptions, they were taken into account when making inferences about stone tool samples.

Table 1
Results of chromatography analysis of sediment and pecked, ground and polished artifacts from CoSC. Fatty acids values are presented as percentages.
Fatty acid

Sample
CoS Sed.

S12 105 2

206

S12 204 1

S12 304 23

S12 304 39

C11:0 e Undecanoic acid


C12:0 e Lauric acid
C13:0 e Tridecanoic acid
C14:0 e Myristic acid
C14:1 e Myristoleic acid
C15:0 e Pentadecanoic acid
C16:0 e Palmitic acid
C16:1 e Palmitoleic acid
C16:4 eHexadecatetraenoic acid
C17:0 e Heptadecanoic acid
C17:1 e Heptadecenoic acid
C18:0 e Stearic acid
C18:1 e Oleic acid
C18:2 e Linoleic acid
C18:3n3 e -Linolenic acid
C18:4 e Stearidonic acid
C19:0 e Nonadecanoic acid
C20:0 e Eicosanoic acid
C20:2 e Eicosadienoic acid
Undeterminated fatty acids

e
e
e
3.036
e
1.612
22.194
1.618
e
1.608
e
32.447
24.569
e
1.976
e
e
e
e
6.613

e
e
e
0.05
e
e
51.30
e
e
e
e
14.1
e
e
e
e
e
e
e
e

e
e
0.186
8.544
0.432
5.147
24.056
12.959
e
2.321
e
22.507
16.804
0.957
e
2.246
e
0.646
e
3.197

e
0.089
e
2.156
0.329
1.58
29.489
3.761
e
2.834
0.927
18.41
32.317
2.011
0.592
0.186
0.111
0.142
1.92
3.145

0.219
0.571
e
7.383
0.372
3.273
29.818
3.89
2.004
2.507
e
22.316
22.12
1.276
e
e
e
e
e
4.249

e
e
e
4.521
0.5
2.031
23.981
2.378
e
2.68
0.616
28.194
22.54
6.848
e
0.788
e
e
4.447
0.477

Table 2
Results of chromatography analysis of sediment and shtail projectile points from CoSC. Fatty acids values are presented as percentages.
Fatty acid

C10:0
C11:0
C12:0
C13:0
C14:0

e
e
e
e
e

Capric acid
Undecanoic acid
Lauric acid
Tridecanoic acid
Myristic acid

Sample
CoS Sed.

S13 905 3

S11 W130 4

S12 404 2

S12 403 2

S12 404 3

S12 404 1

S12 304 1

S12 402 1

S12 406 14

e
e
e
e
3.036

e
e
e
e
0.511

e
e
e
e
2.607

0.015
e
0.297
0.147
2

0.709
e
0.618
0.976
3.546

0.619
0.954
0.446
0.595
1.571

e
e
e
e
29.412

e
e
e
e
0.188

e
e
e
e
4.281

0.033
e
e
0.034
5.327

(continued on next page)

Please cite this article in press as: Mazzia, N., Flegenheimer, N., Detailed fatty acids analysis on lithic tools, Cerro El Sombrero Cima, Argentina,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.027

N. Mazzia, N. Flegenheimer / Quaternary International xxx (2014) 1e13

Table 2 (continued )
Fatty acid

Sample
CoS Sed.

S13 905 3

S11 W130 4

S12 404 2

S12 403 2

S12 404 3

S12 404 1

S12 304 1

S12 402 1

S12 406 14

C14:1 e Myristoleic acid


C15:0 e Pentadecanoic acid
C16:0 e Palmitic acid
C16:1 e Palmitoleic acid
C17:0 e Heptadecanoic acid
C17:1 e Heptadecenoic acid
C18:0 e Stearic acid
C18:1 e Oleic acid
C18:2 e Linoleic acid
C18:3n3 e -Linolenic acid
C18:4 e Stearidonic acid
C19:0 e Nonadecanoic acid
C20:0 e Eicosanoic acid
C20:2 e Eicosadienoic acid
C21:1 e Heneicosenoic acid
C21:2 e Heneicosadienoic acid
C22:2 e Docosadienoic acid
Undeterminated fatty acids

e
1.612
22.194
1.618
1.608
e
32.447
24.569
e
1.976
e
e
e
e
e
e
e
6.613

e
0.762
13.039
0.883
1.703
0.403
23.655
17.314
3.609
0.302
0.731
e
0.445
9.804
0.364
8.586
10.217
7.669

0.195
1.205
19.318
3.274
6.368
0.6
29.709
22.333
3.378
e
e
e
e
4.767
e
e
5.993
0.251

0.217
1.423
22.486
1.847
2.34
0.561
e
22.422
3.446
0.371
0.912
0.192
0.396
3.391
0.134
e
5.118
32.293

0.483
1.278
22.168
2.106
2.781
0.654
32.641
24.896
2.427
0.648
0.388
0.306
0.309
e
e
e
e
3.066

0.15
1.074
17.034
1.477
2.604
0.672
28.994
18.992
3.148
0.685
0.653
0.335
e
8.29
0.453
e
7.287
3.966

e
7.143
32.773
9.034
e
e
14.076
7.563
e
e
e
e
e
e
e
e
e
e

e
e
14.54
0.782
2.503
e
43.53
29.84
3.6
e
e
e
0.735
4.281
e
e
e
e

0.532
1.472
23.192
2.004
2.596
0.515
e
25.46
3.512
0.492
1.327
0.195
0.428
2.47
e
e
2.547
28.976

0.710
1.425
24.059
2.181
2.647
0.547
27.08
27.249
1.875
0.462
0.662
0.188
0.29
1.009
e
e
0.786
3.298

5.1. Pecked and ground lithic artifacts


Table 1 lists the results obtained from the analysis of pecked
and ground lithic artifacts. It does not include the decorated discoidal stone (Fig. 3A) as it has already been described in another
paper (Flegenheimer et al., 2013a). The extract obtained from the
decorated surface was too small, and it only consists of three fatty
acids: myristic, palmitic and stearic. The ubiquity of these compounds in nature makes it impossible to infer any use for the
discoidal stone.
The other four pecked and ground artifacts provided more
abundant and heterogeneous lipid compositions than the discoidal
stone, although one comes from the surface collection. It is artifact
CoSC 206, a fragmented small stone sphere (Fig. 3B) manufactured
on a yellow non-identied quartzitic rock. As this object was more
exposed to oxidation processes than artifacts recovered during
excavation, we expected to obtain a sample containing only saturated fatty acids, mainly of short and medium chains. However, we
found polyunsaturated fatty acids, such as stearidonic acid. The
high proportion of palmitoleic acid and the relative percentages of
stearidonic, myristic, palmitic, oleic and linoleic acids are comparable to compositions that characterize various seed oils (U.S.
Testing Company, N/D; Robinson et al., 1991). It is difcult at the
moment to assess the implications of this result.
The extract S12 204 1 comes from another fragmented small
stone sphere (Fig. 3C) that was made of the Balcarce Formation
orthoquartzite. The relationship among the proportions of palmitic,
palmitoleic, oleic, linoleic, stearidonic and eicosanoic acids is
comparable with the composition of terrestrial animal fat (U.S.
Testing Company, N/D; Robinson et al., 1991). This result would
be consistent either with a leather wrap or thong used to tie the
sphere or its use as a weapon: however, this last seems improbable
due to its small size when compared to later regional bola stones
known to have been used for hunting (Vecchi, 2010). Small spheres
are found in other early assemblages in the Southern Cone both as
manuports and as culturally modied objects of varied raw materials (Dillehay, 1997; Miotti et al., 2010; Flegenheimer et al., 2013a).
Although small spheres were part of early assemblages, their
function or signicance is not yet clear.
A possible fragment of discoidal stone (S12 304 23, Fig. 3D),
manufactured on Balcarce Formation orthoquartzite, had a significant lipid record. The presence of hexadecatetraenoic acid, a
polyunsaturated fatty acid was identied: it has been found in the
databases of the compositions of sh and marine mammal lipids,

and is also present in some algae (Patrick et al., 1985; Brenner and
Bernasconi, 1997; Pond et al., 1997; Sengr et al., 2003; Abd El-Baky
et al., 2004; Muhamad and Mohamad, 2012). Therefore, this object
is linked to the use of aquatic animal resources. As the function of
discoidal stones is unknown so far (Flegenheimer et al., 2013a), this
information is relevant if the fragment corresponds to this type of
artifact.
Finally, a fragment of an unidentied polished artifact made of
quartz (S12 304 39, Fig. 3E) was analyzed. It does not correspond to
a small sphere, so, it either is part of a discoidal stone or of a third
unknown type of ground tool. The amount of fat extracted from this
fragment is remarkable, as its largest side is only 2 cm. In this
sample there are relative values of linoleic, stearidonic, and eicosadienoic acids which can be related to the compositions of vegetable oils (U.S. Testing Company, N/D; Robinson et al., 1991; Mazzia,
2010e2011).
5.2. FTPP
The results of the chromatographic analysis of nine fragmented
FTPP are presented in Table 2. From a typological analysis, some of
these specimens were interpreted as recycled or reused artifacts
with a function different from hunting (Flegenheimer et al., 2013b),
and results were expected to reveal a variety of resources. The
recycled points comprise two points that have been retouched,
possibly after breakage, and currently exhibit bifacial scraping
edges, one of them with a graver at the end. Another point has been
reworked into a drill, and a fourth specimen has been used in
strong abrasive work.
All the samples obtained were sufciently abundant for the gas
chromatograph injection. However, this does not mean that the
past use of the projectile points could be identied in all the artifacts: such is the case of sample S12 404 1 (Fig. 4A). It was obtained
from a fragmented and recycled FTPP that preserves the stem and
part of the blade. It was manufactured on a brownish orthoquartzite (SBGO). The extract obtained was small with few fatty
acids that are not diagnostic and that were also detected in the
sedimentary matrix. Thus, it is not possible to infer the use of this
object in the past.
The fragmented FTPP S12 404 2 (Fig. 4B) was manufactured on
pinkish SBGO, and it also still has the stem and part of the blade.
The lipid sample has a composition with a diversity of fatty acids.
There was no record of stearic acid, which is widespread in nature.
Eicosadieonoic, heneicosenoic, and docosadienoic acids point to a

Please cite this article in press as: Mazzia, N., Flegenheimer, N., Detailed fatty acids analysis on lithic tools, Cerro El Sombrero Cima, Argentina,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.027

Fatty acid

C9:0 e Nonanoic acid


C10:0 e Capric acid
C11:0 e Undecanoic acid
C12:0 e Lauric acid
C13:0 e Tridecanoic acid
C14:0 e Myristic acid
C14:1 e Myristoleic acid
C15:0 e Pentadecanoic acid
C16:0 e Palmitic acid
C16:1 e Palmitoleic acid
C16:4 e Hexadecatetraenoic
acid
C17:0 e Heptadecanoic acid
C17:1 e Heptadecenoic acid
C18:0 e Stearic acid
C18:1 e Oleic acid
C18:2 e Linoleic acid
C18:3n3 e -Linolenic acid
C18:4 e Stearidonic acid
C19:0 e Nonadecanoic acid
C20:0 e Eicosanoic acid
C20:1 e Eicosenoic acid
C20:2 e Eicosadienoic acid
C21:1 e Heneicosenoic acid
C22:2 e Docosadienoic acid
Undeterminated fatty acids

Sample
CoS Sed.

S12 106 6

S12 203 38

S12 404 8

S12 405 11

S7 102 1

S12 205 bis5

S13 27 12

S12 203 7

S12 401 10

S12 4 2

S13
904 14

S12 4 12

S12 305 2

S12
4 11

S12 303 3

e
e
e
e
e
3.036
e
1.612
22.194
1.618
e

e
e
e
e
e
5.947
0.811
1.684
23.5
1.966
e

e
0.268
e
0.399
e
6.649
0.664
2.44
32.565
2.547
1.468

e
e
e
e
e
1.598
0.18
0.842
17.336
1.854
e

e
0.16
e
0.345
0.163
1.983
0.222
1.544
22.33
3.212
e

e
e
e
e
e
2.026
0.201
1.281
19.459
2.735
e

e
e
e
0.204
e
4.183
0.292
2.075
26.363
2.798
1.612

0.689
0.611
1.416
0.895
0.794
8.338
1.087
2.061
23.773
2.569
e

e
e
e
e
e
2.669
0.226
1.705
21.825
3.459
e

e
e
e
e
e
0.235
e
0.207
15.203
1.235
e

e
e
e
e
e
2.875
0.279
1.117
21.915
1.878
e

e
e
e
e
e
2.361
e
1.139
20.532
2.912
e

e
e
e
e
e
0.541
e
0.451
18.13
1.59
e

e
e
e
e
e
1.065
0.174
0.746
14.239
1.015
e

e
e
e
0.039
0.027
4.537
0.608
1.428
23.285
2.348
e

e
e
e
0.142
0.125
7.998
0.632
2.629
23.043
6.153
e

1.608
e
32.447
24.569
e
1.976
e
e
e
e
e
e
e
6.613

2.811
0.409
28.108
21.411
2.153
0.351
0.199
0.191
0.322
e
3.842
0.262
3.779
2.256

2.619
0.442
24.149
20.709
3.549
e
0.931
e
e
e
e
e
e
1.232

2.462
0.467
28.945
23.6
3.561
0.364
0.568
e
0.453
e
9.131
e
8.05
0.59

2.683
0.643
24.705
25.376
3.565
0.328
1.122
e
0.383
e
3.806
e
5.275
2.153

2.408
0.638
26.048
21.226
3.312
0.357
0.779
0.246
0.466
e
8.468
0.22
8.08
1.322

2.483
0.734
27.876
25.026
5.048
e
e
e
e
e
e
e
e
1.304

2.209
0.609
25.263
23.387
3.311
e
e
e
e
e
e
e
e
2.986

2.658
0.618
31.746
21.893
2.603
0.254
0.079
e
0.683
e
3.963
e
e
2.982

2.582
0.523
42.45
31.485
3.117
e
e
e
e
e
e
e
e
e

2.566
0.554
35.646
25.18
3.888
e
e
e
e
e
3.611
e
e
e

2.818
0.821
34.955
30.55
3.276
e
e
e
e
e
e
e
e
e

2.264
e
39.098
26.136
4.436
e
1.017
e
e
e
5.612
e
e
e

1.786
0.394
26.497
17.455
2.611
e
0.614
e
e
e
7.22
e
e
7.374

2.485
0.549
26.538
28.715
e
0.502
0.59
0.188
0.3
0.08
1.569
e
e
e

2.689
e
25.051
21.897
1.506
0.189
1
0.228
0.39
e
1.194
e
e
1.617

N. Mazzia, N. Flegenheimer / Quaternary International xxx (2014) 1e13

Please cite this article in press as: Mazzia, N., Flegenheimer, N., Detailed fatty acids analysis on lithic tools, Cerro El Sombrero Cima, Argentina,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.027

Table 3
Results of chromatography analysis of sediment and retouched artifacts from CoSC. Fatty acids values are presented as percentages.

N. Mazzia, N. Flegenheimer / Quaternary International xxx (2014) 1e13

Fig. 3. Pecked, ground and polished artifacts from CoSC. A. S12 105 2; B. 206; C. S12
204 1; D. S12 304 23; E. S12 304 39.

plant contribution to the substances. However, the sample exhibits


more than 32% undetermined compounds, making it impossible to
infer the origin of the organic resources that have remained on it, as
this situation masks the relationship among fatty acids.
Similarly, sample S12 402 1 (Fig. 4C) has almost 29% of undetermined fatty acids and no record of stearic acid. It is a recycled
artifact, possibly originally a projectile point on red and white
SBGO. The edges of this tool exhibit strong abrasion, most probably
due to use. However, according to the results, its use in processing
particular organic resources cannot be inferred from this analysis.
The lipid sample taken from specimen S13 905 3 (Fig. 4D)
contains interesting information. It is a FTPP which has been
recycled as a drill, manufactured on yellowish SBGO, and still has

Fig. 4. Fishtail projectile points (FTPP) from CoSC. A. S12 404 1; B. S12 404 2; C. S12
402 1; D. S13 905 3; E. S11 130 4; F. S12 404 3; G. S12 406 14; H. S12 403 2; I. S13 27 12;
J. S12 303 3.

Fig. 5. Retouched artifacts from CoSC. A. S13 904 14; B. S12 4 2; C. S12 401 10; D. S7 102
1; E. S12 404 8; F. S12 405 11; G. S12 106 6; H. S12 203 7; I. S13 27 12; J. S12 303 3; K.
S12 4 11; L. S12 305 2; M. S12 4 12; N. S12 205bis 5; O. S12 203 38.

the stem and part of the original blade. It exhibits a signicant


proportion of long/very long chain fatty acids, both saturated and
unsaturated. Among them, eicosadienoic, docosadienoic, heneicosenoic, and heneicosadienoic are noteworthy. Eicosadienoic acid
was identied in the experimental samples of wood, mastic, and
plants. Docosadienoic acid was recognized in experimental samples
of wood and plants, and the heneicosenoic and heneicosadienoic
acids were detected only in plant samples (Mazzia, 2010e2011). In
sum, it is probable that the analyzed substances are the result of the
projectile point haft. However, the presence of fatty acids identied
in vegetable oils, but not in wood or mastic, suggests the contribution of other substances once the point was fragmented and
reutilized after recycling.
Sample S11 130 4 (Fig. 4E) comes from a fragmented FTPP made
of quartz, and only the stem has remained. It presents a relatively
high percentage of heptadecenoic acid. This is interpreted as the
result of bacterial action on substances (Robinson et al., 1991;
Buonasera, 2007). The presence of eicosadienoic and docosadienoic acids refers to the projectiles haft and/or the possible re-use of
its edges on vegetal resources. This FTPP does not present macroscopic evidence of recycling.
Two other samples indicate that the projectile points were
probably hafted. Extract S12 404 3 (Fig. 4F) comes from a pinkish
SBGO stem. As discussed previously, relative proportions of eicosadienoic and docosadienoic acids are interpreted as evidence of
hafting, as they were isolated in experimental samples of wood,
mastic, and plants (eicosadienoic), and in experimental samples of
wood and plants (docosadienoic). It is not possible to dene other
resources from the remaining compounds. The same two fatty acids
were detected in sample S12 406 14 (Fig. 4G), although in different
proportions. This sample corresponds to a complete and maintained FTPP worked on yellow SBGO. It is the only complete
analyzed specimen which has not been recycled. It exhibits

Please cite this article in press as: Mazzia, N., Flegenheimer, N., Detailed fatty acids analysis on lithic tools, Cerro El Sombrero Cima, Argentina,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.027

N. Mazzia, N. Flegenheimer / Quaternary International xxx (2014) 1e13

evidence of a possible vegetal haft, and the relationship among the


proportions of palmitic, palmitoleic, oleic, linoleic, stearidonic and
eicosanoic acids is comparable with the composition of terrestrial
animal fat (U.S. Testing Company, N/D; Robinson et al., 1991). This
result is highly consistent with the expectations for a weapon and
implies that this point was probably used after maintenance.
S12 403 2 (Fig. 4H) is a fragmented and recycled FTPP on whitish
SBGO that preserves part of the blade and only a very small section of
the stem. The point has been reworked into a composite tool with a
small graver on the distal edge. No fatty acids related to hafting were
identied, probably due to the small size of the remaining stem. The
relative percentages of fatty acids in this lipid composition indicate
that this specimen was used, but do not allow an unambiguous
inference about the origin of the resources. Similarly, sample S12
304 1 (Fig. 4I) shows a general undetermined composition. However,
this composition includes a signicant proportion of eicosadienoic
acid, possibly related to a vegetal haft. This is consistent with the
object, which is a stem of a fragmented FTPP of chert.
5.3. Other aked tools
The last group of results comes from 15 aked artifacts that
include different morphologies (Table 3): 14 are made of SBGO and
only one is manufactured on silicied limestone. This group includes different manufacturing sequences. Five artifacts only have
unifacial retouch, two have bifacial retouch and eight exhibit bifacial thinning.
Sample S13 904 14 (Fig. 5A) was extracted from a fragmented
large double convergent side scraper. It has relative proportions of
myristic, palmitic, stearic, oleic and linoleic acids comparable to
terrestrial animal fats (U.S. Testing Company, N/D; Robinson et al.,
1991; Mazzia, 2010e2011). Similarly, the extract taken from a
fragmented small irregular biface (S12 4 2, Fig. 5B) exhibits the
same fatty acids in a proportion indicating an animal origin.
However, this biface also presents a relative percentage of eicosadienoic acid that indicates a plant contribution to the sample,
possibly due to a vegetal haft, although no macroscopic indication
of hafting is evident.
In a fragment of an unidentied composite tool (S12 401 10,
Fig. 5C), a fatty acid prole similar to those described for terrestrial
animal fats was registered. However, the proportion of the stearic
acid appears augmented, probably due to the degradation of more
unstable, long chain and unsaturated fatty acids. In consequence, it
is not possible to infer an unequivocal origin of the organic
resources.
Sample S7 102 1 (Fig. 5D) comes from a fragmented side scraper
on a at ake, possibly a bifacial thinning ake. Although the lipid
record shows a variety of fatty acids, only a few can be differentiated from those detected in the sediments. Among them, linoleic,
eicosadienoic, heneicosenoic and docosadienoic acids indicate a
vegetable origin of the substances. The last three fatty acids were
registered in the experimental samples of vegetables (C20:2 in
wood, mastic and vegetables; C21:1 in vegetables and C22:2 in
wood and vegetable samples: Mazzia, 2010e2011). Based on these
evidences, the use of this tool on wood or/and on different vegetable tissues can be inferred: due to its morphology it is unlikely,
though not impossible, that this object was hafted. Likewise,
extract S12 404 8 (Fig. 5E) also has fatty acids that were identied in
the sediments, as well as linoleic, eicosadienoic, and docosadienoic
acids. According to the experimental references, these fatty acids
can be related to wood, mastic or fruits and leaves of different
plants. This fragment corresponds to a small bifacial tool with one
regular edge and another sinuous edge. Acids possibly include
substances from a vegetable haft or/and vegetable oils, as residues
of plant processing.

Sample S12 405 11 (Fig. 5F) comes from a bifacial fragment,


possibly a double convergent side scraper or an asymmetrical point
tip. In this sample, capric and eicosanoic acids were registered at
less than 1%. These fatty acids have a restricted distribution in nature, but they can be found in these proportions both in animal fats
and vegetable oils (Babot et al., 2007; Buonasera, 2007). The presence of linoleic, eicosadienoic and docosadienoic acids is interpreted as indicative of vegetable substances (Mazzia, 2010e2011).
A composite tool, with notches on one edge and a side scraper
on the opposite, manufactured on a white bifacial thinning ake
(S12 106 6, Fig. 5G) presents ambiguous results. Compared with the
sediment sample, this extract exhibits few diagnostic compounds.
The presence of linoleic, eicosadienoic, heneicosenoic, and docosadienoic acids can be considered indicators of a plant origin, but
their relative percentages are small. Therefore, it is difcult to make
accurate inferences about the origin of the substances recovered.
Similarly, sample S12 203 7, coming from a fragmented bifacial
blank (Fig. 5H), presents low relative proportions of linoleic and
eicosadienoic acids but no other clear indicators. Hence, these two
specimens may have been used in the past on indeterminate
organic resources that could include plants.
The sample obtained from a bifacial edge, possibly a side scraper
(S13 27 12, Fig. 5I) shows a heterogeneous fatty acids prole but
with non-diagnostic compounds. However, some peculiarities can
be mentioned. Nonanoic acid is not found in nature, although it
results from the oxidation of other fatty acids such as oleic, linoleic,
and linolenic. It is often described for rancid vegetable oils or
substances that suffered breakdown after discard of the artifact
(Babot et al., 2007; Buonasera, 2007). The undecanoic, tridecanoic,
pentadecanoic and heptadecanoic acids identied result from
bacterial action on substances (Robinson et al., 1991; Buonasera,
2007). Among them, undecanoic and tridecanoic acids are less
frequent in the reference databases, and they were detected only in
the experimental sample that mixed meat and vegetables (Mazzia,
2010e2011).
A composite bifacial tool made of silicied limestone (S12 303 3
Fig. 5J) presents a fatty acids prole interpreted as a mixture of
substances of different origins. This tool is expected to have a long
life history as it was probably manufactured some 400e500 km
from CoSC (Flegenheimer et al., 2003). It possibly is a recycled
bifacial artifact and its morphology allows different functions
(scraper, graver and abrupt edge). The relationship among the
percentages of myristic, palmitic, stearic, oleic, linoleic, linolenic
and stearidonic acids is comparable to the compositions of terrestrial animal fat (Robinson et al. 1991; U.S. Testing Company, Inc.).
However, the presence of eicosadienoic acid was also identied at
less than 2%, and indicates a minor contribution of vegetable oils.
The extract S12 4 11 (Fig. 5K) exhibits similar results including a
small proportion of eicosenoic acid. This extract comes from an
unclassied fragment of artifact manufactured by bifacial thinning.
Another two samples present an important but not diagnostic
lipid record that can be compared to the experimental sample
corresponding to mixed meat and vegetables (Mazzia, 2010e2011).
One of these samples comes from a fragment of a composite tool
which ends in a drill (S12 305 2 Fig. 5L) and the other from a
fragment of a scraper (S12 4 12 Fig. 5M).
Samples S12 205bis 5 (Fig. 5N) and S12 203 38 (Fig. 5O) exhibit
very similar fatty acids proles. The rst one corresponds to a
fragment of a unifacial double convergent side scraper, and the
other comes from a fragment of a biface. Both have hexadecatetraenoic acid, a polyunsaturated fatty acid with a very restricted
distribution in nature. This acid was also detected in a fragment of a
possible discoidal stone. The only references found for this fatty
acid are in compositions of sh and marine mammal lipids, and it is
also present in some algae (Patrick et al., 1985; Brenner and

Please cite this article in press as: Mazzia, N., Flegenheimer, N., Detailed fatty acids analysis on lithic tools, Cerro El Sombrero Cima, Argentina,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.027

10

N. Mazzia, N. Flegenheimer / Quaternary International xxx (2014) 1e13

Bernasconi, 1997; Pond et al., 1997; Sengr et al., 2003; Abd El-Baky
et al., 2004; Muhamad and Mohamad, 2012). The relative proportions of the other fatty acids identied in these archaeological
samples are consistent with the interpretation assigning a marine
origin to the sample.
In summary, twenty-nine lithic objects with different sizes and
morphologies were analyzed by gas chromatography. Twentyseven have lipid samples with heterogeneous fatty acids compositions that allow us to propose their use in the past for processing
organic resources. Only two, a recycled point (S12 404 1) and a
discoidal stone (S12 105 2), gave samples which do not provide
information about their past use.
6. Discussion
As a rst comment, we want to highlight the important preservation of lipid molecules absorbed in lithic artifacts used more
than 10,000 years ago. This preservation occurred in an environment where no other organic remains survived and in stone tools
that were water washed and then stored for more than twenty
years in the lab. This application of a method usually used on other
materials, such as pottery and grinding tools, is therefore very
promising and will be further explored in order to obtain more
accurate interpretations. Although most of the analyzed artifacts
come from stratigraphy, one of the fragmented spheres (206) was
recovered during surface collection. This sample shows good
preservation. The size of the tools has to be emphasized, as they are
smaller than those commonly analyzed: the sampled stone tools
include a fragment measuring only 2 cm (S12 304 39, Fig. 2E). The
sizes of the objects are not proportional to the amount of fatty acids
they retain. For example, an extract from this small fragment has an
abundant and heterogeneous fatty acids prole.
Regarding the absence of results in some cases, it is highly
improbable that the recycled FTPP (S12 404 1, Fig. 3A) was not used:
on the contrary, hafting and other fatty acids would be expected.
However, it did not yield positive results. Therefore, the absence of
results cannot be taken as indicating that the object was not used
on organic remains. This is directly relevant to the interpretation of
the discoidal stone, making it incorrect to assume that it was not
used on organic resources.
Results obtained through this methodology do not necessarily
reect the last activities carried out. Rather, they yield a prole of
the artifacts life history, a signicant difference with results obtained through use wear studies. However, both methodologies
have been used in the region to call the attention to the variety of
resources used. Mainly, they have been relevant to enhance the
importance of the vegetable world, specically of wood, in the daily
life of early hunter-gatherers (Leipus, 2004). As faunal remains are
more frequently recovered, research about the peopling process
has until recently over-emphasized the importance of hunting. The
role of plants in the early Pampean hunter gatherers diet is
understudied because of its scarce or null representation in the
archaeological record. Despite of the lack of direct evidences, gas
chromatography studies, use wear analysis and stable isotopes on
human remains are introducing signicant data. The study of plant
resources utilization in past subsistence strategies involves more
than their use as food. Some other related issues are the specic
technologies used for their exploitation, the knowledge of the
environment, and social meanings linked to their use.
According to current interpretations, most of the tools discarded
at CoSC were previously used and broken in other places (Weitzel,
2010). As this method reects tool history, the assemblage at the
hilltop most probably represents a wider range of resources than
those used specically at this place. It might even constitute a large
sample of the organic materials in use by these early Pampean

societies. However, obtaining a truly representative sample requires including tools from several sites, as a high intersite variability has been registered in the microregion (Mazzia and
Flegenheimer, 2012). This sample includes an important amount
of vegetable resources, among which seeds and vegetable hafts
(wood and mastic) can be distinguished, along with a smaller
proportion of terrestrial animals and a few marine resources.
The FTPP are the group of artifacts where expectations are easier
to discuss. This group has yielded very consistent results. Of the
nine samples analyzed, ve have evidence of hafting (S13 905 3;
S12 130 4; S12 404 3; S12 304 1; S12 406 14); a sixth sample that
comes from a recycled blade (S12 403 2), as expected, does not
show evidence of hafting; and the other three specimens have
unidentiable results. The only complete point (S12 406 14) shows
evidence of terrestrial animal fat, relating it to its use in hunting.
Four recycled FTPP were analyzed: of these, results from one (S12
905 3) refer to its use on vegetable resources, and the others did not
give identiable results.
The group of pecked, ground, and polished stone tools has not
yielded an identiable pattern in the results. However, it is interesting that these artifacts, whose function is still unknown (Jackson
and Mndez, 2007; Hermo et al., 2013, Flegenheimer et al., 2013a;
Nami, 2013), gave identiable results. These are very varied even
within the same typological group: for example, small spheres have
shown seed oil (206) and terrestrial animal fat (S12 204 1) and
fragments have yielded samples assigned to vegetable resources
(S12 304 39) and marine fatty acids (S12 304 23). Interpretations
regarding these artifacts are still highly speculative, as many
questions about their function remain.
The third group includes a greater number of objects: only one is
complete and probably corresponds to a recycled tool on a long
distance rock, the others are fragmented, and some cannot be
classied due to breakage. This third group has produced a variety
of results. Three objects register terrestrial animal fat (S13 904 14,
S12 4 12, S12 401 10), one with a vegetable contribution and
another with some uncertainty. Three other artifacts exhibit
vegetable resources (S7 102 1, S12 404 8, S12 405 11), one of which
might also be related to hafting and another might include other
unidentied resources. The next two objects described (S12 106 6,
S12 203 7) revealed use on indeterminate resources, possibly
including plants. One artifact (S13 27 12) is also indeterminate, but
some fatty acids refer to rancid substances and bacterial action.
Four objects (S12 303 3, S12 4 11, S12 305 2; S12 4 12) have been
used both on animal and plant resources and have mixed fatty
acids. The last two objects (S12 203 38, S12 205bis 5) register use on
marine resources. Analysis shows that all of these artifacts have
been used in the past, but no relation has been identied between
tool morphology and the kind of resource. This lack of standardization of form-function for several tool types has also been
observed on other early assemblages through functional studies
based on microscopic analysis (Leipus, 2004). Also, many artifacts
exhibit traces of more than one resource, as many tools in the
assemblage have been multifunctional.
The identication of marine resources in three different artifacts
merits comment. As mentioned, the Atlantic coast must have been
w120 km from CoSC at the moment of occupation. This distance
probably falls within the range covered by early people in a yearly
round or even during logistical journeys (Kelly, 1995). Furthermore,
as people occupying CoSC participated in interaction networks over
400e500 km towards the northeast (Flegenheimer et al., 2003),
they were bound to know the Atlantic coast. At other early sites,
Paso Otero 5 and Cueva Tixi, lithic coastal raw materials have been
identied (Valverde, 2002; Martnez and Gutirrez, 2011), and FTPP
have been found along the present coast (Flegenheimer and Bayn,
1996; Bonomo, 2005). Information obtained through fatty acids

Please cite this article in press as: Mazzia, N., Flegenheimer, N., Detailed fatty acids analysis on lithic tools, Cerro El Sombrero Cima, Argentina,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.027

N. Mazzia, N. Flegenheimer / Quaternary International xxx (2014) 1e13

analysis is interesting because it gives support to the idea that early


settlers in the Pampean region made use of marine resources and
were familiar with this particular environment.
7. Conclusion
The good preservation of the analyzed substances through time
is remarkable, and they do not show signicant effects of degradation processes. This good preservation is indicated by the great
proportion of samples with polyunsaturated and very long chain
fatty acids. Good preservation of fatty acids was also found in other
four early lithic assemblages in the micro-region (Mazzia, 2010e
2011, 2011, 2013), but we cannot establish the taphonomic conditions that made it possible. However, in all of these cases, micro
cracks produced by aking and natural porosities of the rocks must
have played a key role in protecting lipids from degradation. This
matter should be addressed during future analysis.
The fatty acids proles indicate that early Pampean people used
their lithic tools on a diversity of organic resources including seeds,
plants, terrestrial animals and marine resources. The results also
indicate the existence of vegetable hafts and vegetable mastic used
to hold some artifacts. Furthermore, chromatographic analysis
provides evidence of different environments visited by early
groups, and is useful to think about the distances travelled by these
people.
Although gas chromatography analysis already has a long history in archaeological studies, mainly on pottery and grinding stone
tools, (for example, Patrick et al. 1985; Rottlnder, 1990; Caabate
Guerrero and Snchez Vizcano, 1995; Malainey et al. 1999;
Buonasera, 2007; Evershed, 2008; Costa Angrizani and Constenla,
2010; in Argentina: Gonzlez de Bonaveri and Frre, 2002, 2004;
Babot, 2004; Babot et al., 2007; Babot and Hocsman, 2008; Babot
et al., 2008; Frre et al., 2010; Mazzia, 2011, 2012; Bonomo et al.,
in press) the results are still generalized and imprecise. It is not
possible yet to dene the kind of vegetable or animal identied.
More accurate distinctions require precise regional databases
where native plants and animals are characterized. These are now
being produced and will permit more detail in future interpretations in order to obtain more reliable identications of
resources, tasks, and movements of early people.
Acknowledgements
The authors want to thank Pilar Babot for her teaching and
encouragement, Federico Ponce for discussing ideas, Valeria
Lukezic and Agueda Caro Petersen for help with images and Juan
Flegenheimer for help with nal revision of the manuscript. Special
thanks to lab Materia Hnos. (Mar del Plata) and Ings. Ferrari and
Snchez. This work was funded through grants PICT Bicentenario
2010 No.1517 and PIP 112-201101-00177.
References
balos, A., Espuny, M.J., Bermdez, R.C., Manresa, A., 2003. Aplicacin de la Cromatografa de Gases/Espectrometra de Masas (GC/MS) en la caracterizacin
qumica de los polihidroxialcanoatos de Pseudomonas aeruginosa AT10. Revista
Cubana de Qumica XV (2), 3e10.
Abd El-Baky, H., El Baz, F.K., El-Baroty, G.S., 2004. Production of lipids rich in Omega
3 fatty acids from the halotolerant alga Dunaliella salina. Biotechnology 3 (1),
102e108.
Anonymous, 1993. A&G Tcnica. N 10.
vila, D., 2011. Resultados de los fechados radiocarbnicos del sitio Laguna El Doce,
departamento general Lpez, provincia de Santa Fe. Revista de la Sociedad
Argentina de Antropologa XXXVI, 337e343.
Babot, M. del P., 2004. Tecnologa y utilizacin de artefactos de molienda en el
Noroeste Prehispnico (Ph.D. Thesis). Facultad de Ciencias Naturales e IML,
Universidad Nacional de Tucumn.

11

Babot, M.P., Mazzia, N., Bayn, C., 2007. Procesamiento de recursos en la regin
pampeana bonaerense: aportes del instrumental de molienda de las localidades
arqueolgicas El Guanaco y Cerro La China. In: Bayn, C., Pupio, A.,
Gonzlez, M.I., Flegenheimer, N., Frre, M. (Eds.), Arqueologa en las pampas.
Sociedad Argentina de Antropologa, Buenos Aires, pp. 635e660.
Babot, M.P., Hocsman, S., 2008. Cazadores-recolectores, pastores y agricultores en
un contexto transicional: Antofagasta de la Sierra (Puna Meridional Argentina)
-5500e1500 AP. Paper presented at Taller Pastoreo y Modernidad en los Andes.
San Miguel de Tucumn.
Babot, M.P., Hocsman, S., Cattneo, G.R., 2008. Microfossils for assessing the use as
projectile points or knifes of archaeological artefacts from Quebrada Seca 3 site,
Southern Argentinean Puna (ca. 5000e4500 years BP). In: Abstracts of 7 International Meeting in Phytolith Research (7 IMPR) e 4 Encuentro de Investigaciones Fitolticas del Cono Sur (4 EIF). Mar del Plata, Buenos Aires, p. 60.
Barros, P., Messineo, P., 2004. Identicacin y aprovisionamiento de chert o ftanita
en la cuenca superior del Arroyo Tapalqu. Estudios Atacameos 28, 87e103.
Barros, P., Messineo, P., 2006. Modos de abastecimiento y explotacin de materias
primas lticas en la cuenca del arroyo Tapalqu (Olavarra, provincia de Buenos
Aires, Argentina). Habitus 4 (2), 711e737.
Bayn, C., Flegenheimer, N., Zrate, M., Deschamps, C., 2004. .Y vendrn los
arquelogos en busca de un hueso.Sitio El Guanaco, partido de San Cayetano.
In: Martnez, G., Gutierrez, M., Curtoni, R., Bern, M., Madrid, P. (Eds.), Aproximaciones Arqueolgicas Pampeanas. Teoras, Mtodos y Casos de Aplicacin
Contemporneos. Fac. de Cs. Sociales, UNCPBA, Buenos Aires, pp. 247e258.
Bayn, C., Flegenheimer, N., Valente, M., Pupio, A., 1999. Dime cmo eres y te dir de
dnde vienes: procedencia de rocas cuarcticas en la Regin Pampeana, vol.
XXIV. Relaciones de la Sociedad Argentina de Antropologa, pp. 187e235.
Bayn, C., Manera, T., Politis, G., Aramayo, S., 2011. Following the tracks of the rst
South Americans. Evolution, Education and Outreach 4 (2), 205e217.
Bondia Pons, I., 2007. Estudio del perl de cidos grasos en la evaluacin de la dieta
mediterrnea como patrn de dieta saludable en poblaciones europeas (Ph.D.
Thesis). Facultad de Farmacia, Universidad de Barcelona.
Bonomo, M., 2005. Costeando las llanuras. Arqueologa del litoral martimo pampeano. Sociedad Argentina de Antropologa, Buenos Aires.
Bonomo, M., Colobig, M., Mazzia, N., 2013. Anlisis de residuos orgnicos y
microfsiles silceos de la cuchara de cermica del Cerro Tapera Vzquez
(Parque Nacional Pre-Delta, Argentina). Revista do Museu de Arqueologia E
Etnologia da USP (in press).
Brenner, R.R., Bernasconi, A.M., 1997. Aporte de cidos grasos esenciales de las series
n-6 y n-3 a la dieta humana por pescados comestibles del ro Paran. Medicina
57, 307e314.
Buonasera, T., 2007. Investigating the presence of ancient absorbed organic residues
in groundstone using GCeMS and other analytical techniques: a residue study of
several prehistoric milling tools from central California. Journal of Archaeological Science 34, 1379e1390.
Caabate Guerrero, M.L., Snchez Vizcano, A., 1995. Anlisis de indicadores bioqumicos del contenido de recipientes arqueolgicos. Complutum 6, 281e291.
Cavallotto, J.L., Violante, R., Nami, H., 2002. Late Plesitocene/Holocene Paleogeography and coastal evolution at the mouth of the Ro de la Plata: implications for
dispersal of paleoindian people in South America. Current Research in the
Pleistocene 19, 13e16.
Charters, S., Evershead, R.P., Goad, L.J., Leyden, A., Blinkhorn, P.W., Denham, V., 1993.
Quantication and distribution of lipid in archaeological ceramics: implications
for sampling potsherds for organic residue analysis and the classication of
vessel use. Archeometry 35 (2), 211e223.
Colombo, M., 2013. Aprovisionamiento de materias primas lticas entre los
cazadores-recolectores pampeanos. In: El caso de las canteras arqueolgicas del
sur del sistema serrano de Tandilia. Facultad de Ciencias Naturales y Museo,
Universidad Nacional de La Plata (Ph.D. Thesis).
Colombo, M., Flegenheimer, N., 2013. La eleccin de rocas de colores por los
pobladores tempranos de las sierras de Lobera (Buenos Aires, Argentina).
Nuevas consideraciones desde las canteras. Boletn del Museo de Arte Precolombino de Chile 18 (1), 125e137.
Costa Angrizani, R., Constenla, D., 2010. Sobre yapeps, aembs y cambuchs:
aproximaciones a la funcionalidad de vasijas cermicas a partir de la determinacin de cidos grasos residuales en tiestos recuperados en contextos
arqueolgicos en el sur de Brasil. In: Bern, M., Luna, L., Bonomo, M.,
Montalvo, C., Aranda, C., Carrera Aizpitarte, M. (Eds.), Maml Mapu pasado y
presente desde la arqueologa pampeana, pp. 215e224. Ed. Libros del Espinillo,
Buenos Aires.
Dillehay, T., 1997. Monte Verde: a Late Pleistocene Settlement in Chile. In: The
Archaeological Context and Interpretation, vol. 2. Smithsonian Institution Press,
Washington.
Evershed, R.P., 1993. Biomolecular archaeology and lipids. World Archaeology 25,
74e93.
Evershed, R.P., 2008. Organic residue analysis in archaeology: the archaeological
biomarker revolution. Archaeometry 50 (6), 895e924.
Evershead, R.P., Heron, C., Charters, S., Goad, L.J., 1992. Chemical analysis of organic
residues in ancient pottery: methodological guidelines and applications. In:
White, R., Page, H. (Eds.), Organic Residues in Archaeology: Their Identication
and Analysis. United Kingdom Institute for Conservation, Archaeology Section,
York, pp. 11e24.
Fankhauser, B., 1994. Protein and lipid analysis of food residues. In: Hather, J.G. (Ed.),
Tropical Archaeobotany: Applications and New Developments. Routledge,
London, pp. 227e250.

Please cite this article in press as: Mazzia, N., Flegenheimer, N., Detailed fatty acids analysis on lithic tools, Cerro El Sombrero Cima, Argentina,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.027

12

N. Mazzia, N. Flegenheimer / Quaternary International xxx (2014) 1e13

Feiser, L.F., Feiser, M., 1960. In: Qumica Orgnica. Girjalbo, Mxico.
Fezler, D., 1995. Rhea oil. The Ratite Encyclopedia. Ostrich, Emu, Rhea, pp. 245e250.
Flegenheimer, N., 1991. La Liebre, un sitio de cantera-taller, vol. 2. Boletn del Centro,
pp. 58e64.
Flegenheimer, N., 2003. Cerro El Sombrero: a locality with a view. In: Miotti, L.,
Salemme, M., Flegenheimer, N. (Eds.), Where the South Winds Blow. Ancient
Evidence of Paleo South Americans. Texas A&M University Press, pp. 51e56.
Flegenheimer, N., Bayn, C., 1996. Surface Fells Cave Stemmed points in the
Argentine Pampas. Current Research in the Pleistocene 13, 17e19.
Flegenheimer, N., Bayn, C., Valente, M., Baeza, J., Femenas, J., 2003. Long distance
tool stone transport in Argentine Pampas. Quaternary International 109e110,
49e64.
Flegenheimer, N., Mazzia, N., Babot, M., 2013a. Estudios de detalle sobre una piedra
discoidal pampeana. Intersecciones en Antropologa 14, 499e505.
Flegenheimer, N., Zrate, M., 1993. The Archaeological Record in Pampean Loess
Deposits. Quaternary International 17, 95e100.
Flegenheimer, N., Miotti, L., Mazzia, N., 2013b. Rethinking early objects and landscape in the Southern Cone: shtail point concentrations in the Pampas and
Northern Patagonia. In: Graf, K., Ketron, C., Waters, M. (Eds.), Paleoamerican
Odyssey. Center for the Study of First Americans, Texas, pp. 359e376.
Frre, M., Constenla, D., Bayn, C., Gonzlez, M.I., 2010. Estudios actualsticos sobre
recursos silvestres mediante el empleo de anlisis qumicos. In: Bern, M.,
Luna, L., Bonomo, M., Montalvo, C., Aranda, C., Carrera Aizpitarte, M. (Eds.),
Maml Mapu pasado y presente desde la arqueologa pampeana, pp. 215e226.
Ed. Libros del Espinillo, Buenos Aires.
Gonzlez de Bonaveri, M.I., Frre, M., 2002. Explorando algunos usos prehispnicos
de la alfarera pampeana. In: Mazzanti, D., Bern, M., Oliva, F. (Eds.), Del Mar a
los Salitrales. Diez Mil Aos de Historia Pampeana en el Umbral del Tercer
Milenio. Univ. Nacional de Mar del Plata, Buenos Aires, pp. 31e40.
Gonzlez de Bonaveri, M.I., Frre, M., 2004. Analysis of potsherd residues and vessel
use in hunter-gatherer-sher groups (Pampean Region, Argentina). In: Le
Secrtariat du Congrs (Ed.), Acts of the XIVth UISPP Congress, BAR Series 1270.
University of Lige, Oxford, pp. 27e35.
Guilderson, T.P., Burkle, L., Hemming, S., Peltier, W.R., 2000. Late Pleistocene sea
level variations derived from the Argentine Shelf. Geochemistry, Geophysics,
Geosystems 1. Paper number 2000GC000098.
Gunstone, F.D., Harwood, J.L., Dijkstra, A.J. (Eds.), 2007. The Lipid Handbook, 3rd
Edition. CRC Press, Boca Raton.
Gutirrez, M., Martnez, G., Luchsinger, H., Grill, S., Zucol, A., Hassan, G., Barros, M.P.,
Kaufmann, C., lvarez, M.C., 2011. Paleoenvironments in the Paso Otero locality
during Late PleistoceneeHolocene (Pampean Region, Argentina): an interdisciplinary approach. Quaternary International 245, 37e47.
Hermo, D., Terranova, E., Marchionni, L., Magnin, L., Mosquera, B., Miotti, L., 2013.
Piedras o litos discoidales en norpatagonia: evidencias en la Meseta de
Somuncur (Ro Negro, Argentina). Intersecciones en Antropologa 14, 507e511.
IUPAC, 1997. In: McNaught, A.D., Wilkinson, A. Blackwell (Eds.), Compendium of
Chemical Terminology. Scientic Publications, Oxford. XML on-line corrected
version: http://goldbook.iupac.org (2006) created by M. Nic, J. Jirat, B. Kosata;
updates compiled by A. Jenkins. ISBN 0-9678550-9-8. doi:10.1351/goldbook.
Jackson, D., Mndez, C., 2007. Litos discoidales tempranos en contextos de Patagonia. Magallania 35 (1), 43e52.
Johnson, E., Holliday, V., Martnez, G., Gutirrez, M., Politis, G., 2012. Geochronology
and landscape development along the middle ro Quequn Grande at the Paso
Otero locality, Pampa Interserrana, Argentina. Geoarchaeology. An International
Journal 27, 300e323.
Kelly, R., 1995. Foraging and mobility. In: The Foraging Spectrum. Smithsonian
Institution Press, Washington DC, pp. 111e160.
Leipus, M., 2004. Evidencias del uso sobre madera de artefactos lticos manufacturados por talla en el rea Interserrana: El aporte del anlisis funcional. In:
Martnez, G., Gutierrez, M., Curtoni, R., Bern, M., Madrid, P. (Eds.), Aproximaciones Arqueolgicas Pampeanas. Teoras, Mtodos y Casos de Aplicacin
Contemporneos. UNCPBA, Buenos Aires, pp. 147e168. Fac. de Cs. Sociales.
Leipus, M., 2010. El uso de los conjuntos lticos tempranos de Tandilia (Regin
Pampeana, Argentina): evidencias a partir del anlisis funcional de base
microscpica. In: Abstracts of the Vth. International Symposium Early Man in
America, La Plata, Argentina, pp. 80e81.
Malainey, M.E., 2007. Fatty acid analysis of archaeological residues: procedures and
possibilities. In: Barnard, H., Eerkens, J.W. (Eds.), Theory and Practice in
Archaeological Residue Analysis, BAR International Series 1650. Archeopress,
Oxford, pp. 77e89.
Malainey, M.E., Przybylski, R., Sherriff, B.L., 1999. The fatty acid composition of
native food plants and animals of Western Canada. Journal of Archaeological
Science 26, 83e94.
Martnez, G., Gutirrez, M., 2004. Tendencias en la explotacin humana de la fauna
durante el Pleistoceno nal y Holoceno en la Regin Pampeana (Argentina). In:
Mengoni Goalons, G. (Ed.), Zooarchaeology of South Amrica, B.A.R. (International Series) 1298, pp. 81e98. Oxford.
Martnez, G., Gutirrez, M., 2011. Paso Otero 5: a summary of the interdisciplinary
lines of evidence for reconstructing early human occupation and paleoenvironment in the Pampean Region, Argentina. In: Vialou, D. (Ed.), Peuplements et
Prhistoire en Ameriques. Museum National dHistoire Naturelle, UMR, France,
pp. 271e284.
Martnez, G., Gutirrez, M., Tonni, E., 2013. Paleoenvironments and faunal extinctions: analysis of the archaeological assemblages at Paso Otero locality

(Argentina) during the Late PleistoceneeEarly Holocene. Quaternary International 299, 53e63.
Mazzanti, D.L., 2003. The PleistoceneeHolocene stratigraphic record from early
archaeological sites in caves and rockshelters of Eastern Tandilia, Pampean
Region, Argentina. In: Miotti, L., Salemme, M., Flegenheimer, N. (Eds.), Where
the South Winds Blow. Ancient Evidence of Paleo South Americans. Texas A&M
University Press, pp. 57e61.
Mazzanti, D., Colobig, M.M., Zucol, A., Martnez, G., Porto Lpez, J., Brea, M.,
Passeggi, E., Soria, J.L., Quintana, C., Puente, V., 2010. Investigaciones arqueolgicas en el sitio 1 de la localidad Lobera 1. In: Bern, M., Luna, L., Bonomo, M.,
Montalvo, C., Aranda, C., Carrera Aizpitarte, M. (Eds.), Maml Mapu pasado y
presente desde la arqueologa pampeana, pp. 99e114. Ed. Libros del Espinillo,
Buenos Aires.
Mazzanti, D., Martnez, G., Quintana, C., 2012. Early settlements in Eastern Tandilia,
Buenos Aires province, Argentina: archaeological contexts and site-formation
processes. In: Miotti, L., Salemme, M., Flegenheimer, N., Goebel, T. (Eds.), Current Research in the Pleistocene, Special Edition: Southbound, the Late Pleistocene Peopling of Latin America, pp. 99e104.
Mazzia, N., 2010e2011. Lugares y paisajes de cazadores-recolectores en la pampa
bonaerense: cambios y continuidades durante el Pleistoceno nal-Holoceno
(PhD. Thesis). Facultad de Ciencias Naturales y Museo, Universidad Nacional de
La Plata.
Mazzia, N., 2011. El Ajarafe: Un espacio serrano ocupado efmeramente en diferentes momentos del Holoceno (Tandilia, Provincia de Buenos Aires). Revista del
Museo de Antropologa 4, 33e46.
Mazzia, N., 2013. Cueva Zoro: nuevas evidencias sobre pobladores tempranos en el
sector centro oriental de Tandilla. Intersecciones en Antropologa 14, 93e106.
Mazzia, N., Flegenheimer, N., 2012. Early settlers and their places in the Tandilia Range
(Pampean region, Argentina). In: Miotti, L., Salemme, M., Flegenheimer, N.,
Goebel, T. (Eds.), Current Research in the Pleistocene, Special Edition: Southbound, the Late Pleistocene Peopling of Latin America, pp. 105e110.
Messineo, P.G., Politis, G.G., 2009. New radiocarbon dates from the Campo Laborde
site (Pampean Region, Argentina) support the Holocene survival of giant
ground sloth and glyptodonts. Current Research in the Pleistocene 26, 5e9.
Miotti, L., Hermo, D., Terranova, E., 2010. Fishtail points: rst evidence of LatePleistocene Hunter Gatherers in Somuncur Plateau (Ro Negro province,
Argentina). Current Research in the Pleistocene 27, 22e24.
Muhamad, N.A., Mohamad, J., 2012. Fatty acids composition of selected Malaysian
shes. Sains Malaysiana 41 (1), 81e94.
Nami, H., 2013. Archaeology, Paleoindian research and lithic technology in the
Middle Negro River, Central Uruguay. Archaeological Discovery 1 (1), 1e22.
Paez, M., Zrate, M., Mancini, M., Prieto, A., 2003. Paleoenvironments during the
PlesitoceneeHolocene Transition in Southern South America, Argentina. In:
Miotti, L., Salemme, M., Flegenheimer, N. (Eds.), Where the South Winds Blow.
Ancient Evidence of Paleo South Americans. Texas A&M University Press,
pp. 121e125.
Parker, G., Violante, R.A., Paterlini, C.M., Costa, I., Marcolini, S., Cavallotto, J., 2008.
Las secuencias depositacionales del Plioceno-Cuaternario en la plataforma
submarina adyacente al litoral del este bonaerense. Latin American Journal of
Sedimentology and Basin Analysis 15 (2), 105e124.
Patrick, M., de Koning, A.J., Smith, A.B., 1985. Gas liquid chromatographic analysis of
fatty acids in food residues from ceramics found in the Southwestern cape,
South Africa. Archeometry 27 (2), 231e236.
Politis, G.G., 2008. The Pampas and Campos of South America. In: Silverman, H.,
Isbell, W.H. (Eds.), Handbook of South American Archaeology. Springer, New
York, pp. 235e260.
Politis, G., Barrientos, G., Stafford, T., 2011. Revisiting Ameghino: New 14C Dates from
Ancient Human Skeletons from the Argentine Pampas. In: Peuplements et
prehistoire en Ameriques, pp. 43e53.
Ponce, F., Rabassa, J., Coronato, A., Borromei, A.M., 2011. Palaeogeographical evolution of the Atlantic coast of Pampa and Patagonia from the last glacial
maximum to the Middle Holocene. Biological Journal of the Linnean Society
103, 363e379.
Pond, D., Dixon, D., Bell, M., Fallick, A., Sargent, J., 1997. Occurrence of 16:2 (n-4) and
18:2 (n-4) fatty acids in the lipids of the hydrothermal vent Shrimps Rimicaris
Exoculata and Alvinocaris Markensis: Nutritional and Trophic Implications. In:
Marine Ecology Progress Series 156, pp. 167e174.
Robinson, D., Calvo Rebollar, M., Sevillano Calvo, E., 1991. Bioqumica y valor
nutritivo de los alimentos. Acribia, Espaa.
Rottlnder, R.C.A., 1990. Lipid Analysis in the Identication of Vessel Contents. In:
MASCA Research Papers in Science and Archaeology, vol. 7, pp. 37e40.
Sengr, F., zden, ., Erkan, N., Tter, M., Ayse Aksoy, H., 2003. Fatty acid compositions of athead grey mullet (Mugil cephalus L., 1758) llet, raw and beeswaxed caviar oils. Turkish Journal of Fisheries and Aquatic Sciences 3, 93e96.
Tonni, E.P., Prado, J.L., Menegaz, A.N., Salemme, M.C., 1985. La Unidad Mamfero
(fauna) Lujanense. Proyeccin de la estratigrafa mamaliana al Cuaternario de la
Regin Pampeana. Ameghiniana 22 (3e4), 255e261.
U. S. Testing Company, INC. N/D. Chemical and Physical Tables. Tables and Data from
United States Testing Company, Inc. New Jersey, 28e29.
Valverde, F., 2002. Variabilidad de Recursos Lticos en dos sitios Paleoindios de las
Sierras de Tandilia Oriental, provincia de Buenos Aires. In: Mazzanti, D.,
Bern, M., Oliva, F. (Eds.), Del Mar a los Salitrales. Diez Mil Aos de Historia
Pampeana en el Umbral del Tercer Milenio. Univ. Nacional de Mar del Plata,
Buenos Aires, pp. 281e287.

Please cite this article in press as: Mazzia, N., Flegenheimer, N., Detailed fatty acids analysis on lithic tools, Cerro El Sombrero Cima, Argentina,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.027

N. Mazzia, N. Flegenheimer / Quaternary International xxx (2014) 1e13


Vecchi, R., 2010. Bolas de boleadora en los grupos cazadores-recolectores de la
Pampa bonaerense (PhD Thesis). Facultad de Filosofa y Letras, Universidad de
Buenos Aires.
Weitzel, C., 2010. El estudio de los artefactos formatizados fracturados. Contribucin
a la comprensin del registro arqueolgico y la actividad humana (PhD Thesis).
Facultad de Filosofa y Letras, Universidad de Buenos Aires.
Weitzel, C., 2012. Broken Stone tools from Cerro El Sombrero Cima (Tandilia Range,
Argentina). In: Miotti, L., Salemme, M., Flegenheimer, N., Goebel, T. (Eds.),
Current Research in the Pleistocene, Special Edition: Southbound, the Late
Pleistocene Peopling of Latin America, pp. 111e115.
Zrate, M., Flegenheimer, N., 1991. Geoarchaeology of Cerro La China Locality
(Buenos Aires, Argentina): site 2 and site 3. Geoarchaeology: An Internacional
Journal 6 (3), 273e294.
Zrate, M., Rabassa, J., 2005. Geomorfologa de la provincial de Buenos Aires. In: de
Barrio, R., Etcheverry, R., Caball, M., Llambas, E. (Eds.), Geologa y Recursos

13

Minerales de la Provincia de Buenos Aires. Relatorio del XVI Congreso Geolgico


Argentino. UNLP, La Plata, pp. 119e138.
Zrate, M., Camilin, C., Iasi, R., 1993. Quaternary stratigraphy and soil development
on granitoid rocks at Cerro El Sombrero, Tandilia Range, Argentina. In:
Rabassa, J., Salemme, M. (Eds.), Quaternary of South America and Antarctic
Peninsula, vol. 8. A. A. Balkema, Rotterdam, pp. 71e84.
Zrate, M., Gonzlez de Bonaveri, M.I., Flegenheimer, N., Bayn, C., 2000/2002.
Cuadernos del Instituto Nacional de Antropologa y Pensamiento Latinoamericano. Sitios arqueolgicos someros: el concepto de sitio en estratigrafa y
sitio de supercie, vol. 19, pp. 635e653.
Zrate, M., Kemp, R., Espinosa, M., Ferrero, L., 2000. Pedosedimentary and palaeoenvironmental signicance of a Holocene alluvial sequence in the southern
Pampas, Argentina. The Holocene 10 (4), 481e488.

Please cite this article in press as: Mazzia, N., Flegenheimer, N., Detailed fatty acids analysis on lithic tools, Cerro El Sombrero Cima, Argentina,
Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.027

Quaternary International xxx (2014) 1e19

Contents lists available at ScienceDirect

Quaternary International
journal homepage: www.elsevier.com/locate/quaint

Evidence of the earliest humans in the Southern Deseado Massif


(Patagonia, Argentina), Mylodontidae, and changes in water
availability
G.A. Brook a, *, N.V. Franco b, P. Ambrstolo c, M.V. Mancini d, L. Wang a, P.M. Fernandez e
a

University of Georgia, Department of Geography, Athens, GA 30602, USA


CONICET, University of Buenos Aires, Argentina
CONICET, University of La Plata, Argentina
d
National University of Mar del Plata, Argentina
e
CONICET-INAPL, University of Buenos Aires, Argentina
b
c

a r t i c l e i n f o

a b s t r a c t

Article history:
Available online xxx

Mylodontidae bones from La Gruta 3 rockshelter, which date to 11,077e10,571 cal BP (9560  30
e9470  30 14C BP) and 9539e9466 cal BP (8540  30 14C BP), indicate that the extinct giant ground
sloth was in the area after it was rst occupied by humans during the late Pleistocene at 12,799
e12,049 cal BP (10,845  61e10,477  56 14C BP). Sediment characteristics at La Gruta 1 and 3 rockshelters (LG1 and LG3) suggest that conditions were wetter during major periods of human occupation
and this is supported by pollen data. Lacustrine silts and clays in La Barda, and La Gruta Lagoons 1 and 3,
provide evidence of an arid interval prior to about 6500 cal BP (5690  35 14C BP) followed by wetter
conditions. This may explain why there is no evidence of humans between ca. 7760 and 5583 cal BP
(7500  250 and 4770  25 14C BP) either at La Gruta or at La Martita and Viuda Quenzana, which are ca.
25 km away. There is considerable evidence for occupation at Viuda Quenzana after 5581 cal BP and
scanty evidence for occupation at La Gruta around 3800 cal BP with more abundant evidence after
1880 cal BP. In the last 1500 years, six radiocarbon ages show that humans occupied LG1 and LG3 before
(1372e1271 cal BP) and after (539e156 cal BP), but not during, the Medieval Climate Anomaly, which
may have been a time of increased aridity in the area. The ndings at La Gruta show that Mylodontidae
was probably present in the southern Deseado Massif after the rst humans arrived but data from
southern Patagonia show that it became extinct soon afterwards.
2014 Elsevier Ltd and INQUA. All rights reserved.

Keywords:
Early humans
Mylodontidae
Patagonia
Caves
Paleoclimate

1. Introduction
Water is an important resource in arid and semiarid environments for hunter-gatherer populations (e.g. Smith et al., 2005;
Veth, 2005). Early human occupation of Patagonia has been
linked to periods with more available water (e.g. Heusser and
Streeter, 1980; Heusser and Rabassa, 1987; Heusser, 1995;
Coronato et al., 1999; Paez et al., 1999; Miotti and Salemme,
2003; Brook et al., 2013).

* Corresponding author.
E-mail addresses: gabrook@uga.edu (G.A. Brook), nvfranco2008@gmail.com (N.
V. Franco), pambrustolo@hotmail.com (P. Ambrstolo), mvmancin@mdp.edu.ar
(M.V. Mancini), njulixinwang@gmail.com (L. Wang), pablomarcelofernand@gmail.
com (P.M. Fernandez).

The late glacial interval was characterized by substantial


changes in effective moisture, related to stepwise changes primarily in temperature. Vegetation throughout southernmost
Patagonia, during the PleistoceneeHolocene transition, in Andean
and extra-Andean regions, was steppe and heath with substantial
open ground with precipitation lower than today (e.g. VillaMartnez and Moreno, 2007; Tonello et al., 2009; Bamonte and
Mancini, 2011). After 11,000 cal BP, the slight but continuous
development of the forest, shown by the increase in Nothofagus
values together with the presence of the grass steppe, reect the
establishment of the forest-steppe ecotone (Mancini, 2009; Sottile
et al., 2012). Signicant changes in the vegetation occurred
throughout Patagonia during the early Holocene (11,500e
8000 cal BP) (Bamonte and Mancini, 2011; Sottile et al., 2012; De
Porras et al., 2012). In the Central Deseado Massif in South Patagonia, and in nearby areas, there are fewer ages indicating human

http://dx.doi.org/10.1016/j.quaint.2014.04.022
1040-6182/ 2014 Elsevier Ltd and INQUA. All rights reserved.

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

occupation during arid periods (Brook et al., 2013), although there


is still no agreement as to how aridity affected human populations
(e.g. Miotti and Salemme, 2003; Borrero, 2012). In fact, a signicant
number of ages documenting the initial human exploration of
Patagonia have come from sites in the Deseado Massif, with more
recent ages being obtained from adjacent areas and areas closer to
the Andean ranges (e.g. Miotti and Salemme, 2003, 2004; Paunero
et al., 2007; Paunero, 2009), while others have been obtained from
the Magellan Strait region (e.g. Massone, 1987, 1996; Mengoni,
1987; Nami, 1985-86; Nami and Nakamura, 1995).
This paper presents evidence showing that the rst huntergatherers in the southern Deseado Massif shared the environment with Lama guanicoe (guanaco) and probably Mylodontidae at
a time when water may have been more abundant than today.
Although we will focus on the latest information from La Gruta
area, information from areas within 25 km of La Gruta with very
variable environmental conditions today, including La Martita Cave
4 (Aguerre, 2003), Viuda Quenzana 8 (Franco et al., 2013) and El
Verano Cave 1 (Durn et al., 2003), will be discussed also, as each of
these areas would have presented different problems and different
opportunities for hunter-gatherers (Figs. 1 and 2).
2. The southern Deseado Massif
The area around La Gruta is an ancient volcanic landscape, with
abundant natural depressions frequently occupied by shallow lagoons, water levels depending on rainfall. Some lagoons are semipermanent, while others are markedly seasonal, the difference
depending largely on the size of the lagoon catchment. A few lagoons with large catchments even have small streams owing into
them. Almost all of the lagoons in the area show evidence of water
levels having been higher in the past, with ancient beach ridges
well above present water levels and thick deposits of silt/clay in the

oors attesting to slow sedimentation in water. Caves and rockshelters in ignimbrites (Baha Laura Formation) or fossiliferous
sandstones of the Monte Len Formation (Panza and Marin, 1998)
are relatively rare but where they occur they frequently preserve
important archaeological and paleontological deposits.
In contrast, the Viuda Quenzana and La Martita areas have
abundant caves, seasonal streams, occasional perennial streams,
and springs are more common than at La Gruta (Fig. 1). Highquality siliceous rocks are more abundant at La Martita and Viuda
Quenzana than La Gruta (Franco et al., 2012), due to differences in
bedrock geology (Panza and Marin, 1998) and the localized occurrence of siliceous rocks deposited by ancient hot springs containing
Fe and Mn (Claudio Iglesias, pers. comm. 2014).
3. Rockshelters at La Gruta
Three rockshelters have been excavated at La Gruta since 2007.
These are in cliffs around the margins of two lagoons (Fig. 2a). La
Gruta 1 is a shallow cave in silicied ignimbrite (Claudio Iglesias,
pers. comm. 2014) overlooking a small lagoon about 6 m below
(Fig. 3a). In contrast, La Gruta 2 and La Gruta 3, which are about
50 m apart, have formed in a fossiliferous sandstone cliff along the
southern margin of a large basin (Fig. 3b). After modest rains, two
shallow lagoons occupy the topographically lowest sections of this
basin but after heavy rains these water bodies combine into a single
lake that lls the entire depression. The oor of La Gruta 2 is about
3 m above the adjacent lagoon oor, while the oor near the back
wall of La Gruta 3 is only about 1.5 m above it. As recently as 1983
the lagoon at La Gruta 2 and La Gruta 3 lled with water to the
extent that La Gruta 3 was ooded. Flotsam along the eastern,
downwind margin of the basin delimits the elevation reached by
the oodwaters and shows that the entire basin was lled by a lake
up to ca. 4 m deep (Figs. 11b and 12a and b). Excavations at the three

Fig. 1. Locations of Southern Deseado sites mentioned in the text. LG: La Gruta; VQ: Viuda Quenzana; EV: El Verano; LMrt: La Martita.

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

Fig. 2. Google Earth satellite images showing La Gruta 1, 2, and 3 rockshelters (triangles), La Gruta Lagoon 1 and 2 sediment sampling locations (circles) LAG1 and LAG2
(A), and La Barda Lagoon sampling location EB1 (B).

La Gruta rockshelters have provided important information,


including bones of extinct animals and evidence of when the rst
humans entered this region of Patagonia, and what conditions were
like at that time.
Munsell colors of samples from sediment units exposed by the
excavations were determined on dry samples using the U.S.
Department of Agriculture soil manual. Samples were then sieved
through a 2 mm sieve. Particles with diameters larger than 2 mm
were weighed and bagged separately, while those smaller than
2 mm were analyzed for grain-size distribution. In this analysis
about 5e10 g of sediment was pretreated using 30% hydrogen
peroxide to remove any organic material and then 50 g/L sodium
metaphosphate solution was added to disaggregate clay agglomerations. The proportions of clay, silt, and sand in each sample were
measured and calculated using the pipette method of Indorante
et al. (1990). The sand portion was dried and sieved at one-phi
intervals. Folk and Ward (1957) descriptive statistics of grain-size
distribution were calculated using GRADISTAT 4.0 (Blott and Pye,
2001) and granulometric histograms showing the particle-size
distribution were constructed to emphasize the different size
modes in each unit (Figs. 5 and 8). In addition, subsamples of the
<2 mm fraction of each sample were subjected to dilute double
acid extraction of easily-soluble elements using the Mehlich 1
method (Mehlich, 1953). The M1 extracting reagent is a mixture of
0.05 N HCl and 0.025 N H2SO4. Extracted elements were measured
using EPA Method 6010C (U.S. EPA, 2014) on the Inductively
Coupled Plasma-Optical Emission Spectrometer (ICP-OES) at the
Center for Applied Isotope Studies, University of Georgia (Figs. 6
and 9).
Palynological information on past vegetation was obtained from
archaeological excavations at La Gruta and from other archaeological sites in the Deseado Massif (La Martita Cave 4, Los Toldos
and La Mara archaeological localities) as well as from mallines to
the west such as La Tercera (Bamonte and Mancini, 2011; Brook
et al., 2013). Vegetation characteristics and paleoclimatic interpretations of the pollen data are based on present-day pollene
vegetation relationships which indicate that shrubs (Asteraceae
subf. Asteroideae) and dwarf shrubs (Ephedra, Nassauvia) record
drier conditions, while higher frequencies of grasses and herbs are
related to wetter conditions (Mancini, 1998; Mancini et al., 2012).

Fig. 3. La Gruta rockshelters. La Gruta 1 in silicied ignimbrite overlooking relict beach ridges in the adjacent lagoon that record high water levels (a). La Gruta 2 and La Gruta 3
rockshelters in a sandstone cliff along the southern margin of a large complex lagoon 2.5 km from La Gruta 1 (b). La Gruta 2 is to the right of the introduced trees in (b) and La Gruta
3 is just to the left of the trees. The photographs were taken in March 2013 after heavy rains that ponded water in the lagoonal depressions.

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

Fig. 4. Simplied stratigraphic sections of the four walls of the La Gruta 1 excavation showing pollen, sediment and radiocarbon sample locations as well as Units A-C (modied
after Mancini et al., 2013). Charcoal samples were collected during excavation so equivalent locations in the prole were determined by extrapolation.

New uncalibrated AMS radiocarbon ages reported here are in


radiocarbon years BP or years before 1950 AD, and were calculated
using a 14C half-life of 5568 years (Tables 1 and 2). Errors are one
standard deviation and reect both statistical and experimental

errors. Ages were corrected for isotopic fractionation to a d13C value of


25&. These and previously published radiocarbon ages in 14C BP
were calibrated at the 2s probability level in calendar years BP (cal BP)
using CALIB 7.0 (Stuiver and Reimer, 1993) and the Southern Hemisphere (SHcal13) atmospheric calibration curve of Hogg et al. (2013).
3.1. La Gruta 1
A 1  1 m test pit excavated in the oor of La Gruta 1 rockshelter was started in 2007 and reached bedrock in 2008. The
excavation exposed three stratigraphic units AeC, two of them (A
and B) showing lithologic variations within the unit (Figs. 4e6;
Table 3). Three sediment samples from basal Unit A, resting on
bedrock, show it to be yellowish brown muddy sandy gravel, with

Fig. 5. Variations in sediment texture between units at La Gruta 1 rockshelter. For


clarity, histograms for adjacent units are shaded differently. VC very coarse;
C coarse; M medium; F ne, and VF very ne. See Fig. 4 for locations in the
prole.

Fig. 6. Variations in elemental concentrations in La Gruta 1 sediment units. Values are


in parts per million (ppm) for particles < 2 mm diameter. Note the different scales in
the upper and lower diagrams.

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

a Folk and Ward (1957) graphic mean grain size of medium sand
(two samples) and ne sand (one sample). All samples are poorly
sorted, very ne skewed and platykurtic (see Table 3 for denitions). The three samples consisted of 32e40% gravel, 33e40%
sand, 15e20% silt, and 8e10% clay (27e29% silt clay). The
overlying Unit B varied laterally with sample LG1-8 being similar
to sediments in Unit A (brownish yellow, muddy sandy gravel,
mean in medium sand range), suggesting it may have been
reworked from the underlying layer. It consists of 40% gravel, 38%
sand, 15% silt, and 8% clay (22% silt clay) and is very poorly
sorted, very ne skewed, and mesokurtic. In contrast, samples
LG1-3 and LG1-4 are somewhat ner being dark yellowish brown
gravelly muddy sand with a mean grain size in the ne sand range.

Both samples are very poorly sorted and very ne skewed; LG1-3
has a platykurtic grain size distribution and LG1-4 a very platykurtic distribution. Gravel ranges from 20 to 26%, sand from 41 to
53%, silt 18 to 20%, and clay 9 to 13% (silt clay is 27e33%).
Therefore, Unit B is ner with a higher content of silt and clay and
less gravel than Unit A. The upper Unit C is somewhat similar to
Unit A; samples from the charcoal-rich layer and from a layer of
ash are dark grey (because of the charcoal and dung mixed with
the sediments) muddy sandy gravel with a mean grain size of
coarse sand, very poorly sorted, and very ne skewed, with kurtosis being mesokurtic (LG1-5) or leptokurtic (LG1-6). Gravel
ranges from 48 to 50%, sand from 32 to 33%, silt 13 to 14%, and clay
4e5% (silt clay is 18e19%).

Table 1
Radiocarbon ages from La Gruta and nearby areas. La Gruta 1, 2 and 3 data are shaded differently for greater clarity.

Site***

Lab ID

Material

Radiocarbon
Age (14C BP)
32,650 140
24,410 35

2 Radiocarbon
Age (cal BP)
36,934-36,136
28,654-28,211

this paper
this paper

10,845 61
10,840 62
10,790 30
10,720 30

12,799-12,653
12,801-12,650
12,730-12,663
12,711-12,562

Franco et al. 2010


Franco et al. 2010
Franco et al. 2010
this paper

10,656 54
10,477 56
9560 30

12,692-12,434
12,549-12,049
11,077-10,678

Franco et al. 2010


Franco et al. 2010
this paper

9470 30
8540 30

10,748-10,571
9539-9466

this paper
this paper

8090 30
7560 30
8050 90
7940 260

9029-8774
8403-8208
9121-8599
9429-8207

Mancini et al. 2013


Franco et al. 2013
Aguerre 2003
Aguerre 2003

charcoal

8960 140

10,294-9551

Durn et al. 2003

charcoal
guanaco bone*h
guanaco bone*h
charcoal
charcoal
charcoal
charcoal

7500 250
4770 25
4740 25
4520 50
4475 95
3487 38
1888 39

8972-7760
5583-5325
5578-5320
5307-4891
5311-4846
3832-3594
1881-1704
(AD 69-246)
1826-1589
(AD 124-361)
1372-1271
(AD 578-679)
539-503
(AD 1411-1447)
493-327
(AD 1457-1623)
491-324
(AD 1459-1626)
438-156
(AD 1512-1794)
438-156
(AD 1512-1794)

Durn et al. 2003


Franco et al. 2013
Franco et al. 2013
Aguerre 1982
Aguerre 1982
Franco et al. 2010
Franco et al. 2010

La Gruta 3(1)
La Gruta 3
Entrance Pit (a)
La Gruta 1(a)
La Gruta 1(b)
La Gruta 1(c)
La Gruta 3(2)

UGAMS-12427
UGAMS-15124

La Gruta 1(d)
La Gruta 1(e)
La Gruta 3-523/1(3)

AA-76792
AA-84225
UGAMS-13611

La Gruta 3-523/2(4)
La Gruta 3-516(5)

UGAMS-15766
UGAMS 15765

La Gruta 1(f)
La Gruta 2
La Martita Cave 4
La Martita Cave 4

El Verano Cave 1
Viuda Quenzana 8
Viuda Quenzana 8
La Martita Cave 4
La Martita Cave 4
La Gruta 1(g)
La Gruta 1(h)

UGAMS-7540
UGAMS-9113
CsIC-506+
Teledyne Isotopes
I.11, 903
Teledyne Isotopes
I.13, 797-No 1
INGEIS 2854
UGAMS-9111
UGAMS-9112
CsIC-505+
I-11904
AA-84226
AA-83474

La Gruta 1(j)

AA-83476

charcoal

1829 47

La Gruta 1(i)

AA-83475

charcoal

1452 38

La Gruta 3(6)

UGAMS-13609

charcoal

530 20

La Gruta 1(k)

UGAMS-7541

charcoal

400 20

La Gruta 3(7)

UGAMS-13610

charcoal

390 20

La Gruta 3(8)

UGAMS-12430

guanaco bone*h

290 20

La Gruta 3(9)

UGAMS-12429

guanaco bone*h

290 20

El Verano Cave 1

AA-84224
AA-84223
UGAMS-7538
UGAMS- 12428

puma rib*
unidentified
bone**
charcoal
charcoal
charcoal
guanaco phalanx
bone*
charcoal
charcoal
Mylodontidae
vertebra**
Mylodontidae**
Mylodontidae
vertebra**
charcoal
guanaco bone*h
charcoal
charcoal

Reference

Franco et al. 2010


Franco et al. 2010
this paper
this paper
this paper
this paper
this paper

* bone collagen age; ** bone bioapatite age; h bone shows evidence of human action.
*** LG1 = La Gruta 1, LG2 = La Gruta 2, LG3 = La Gruta 3. Radiocarbon sample locations (in parentheses) are shown in Figs. 4 (LG1) and 7
(LG3).
+
Laboratorio de Geocronologa, Instituto de Qumica Fsica, Rocasolano.

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

Fig. 7. Stratigraphic prole of the LG3 Main Excavation showing Units 1 through 6 and sediment, pollen, and radiocarbon dating sample locations, and a diagrammatic prole of
Entrance Pit near the drip line of the shelter showing Units EA to EC. Bone and charcoal samples were collected during excavation so equivalent locations in the prole were
determined by extrapolation.

Elemental concentrations at LG1 show trends within and between units (Fig. 6). The basal sediments of Unit A (A-7 and A-2) have
higher values of Al and Fe (1903 and 2625 ppm) and relatively low
values of the more soluble elements Ca, Mg, Na and K (231, 21, 39, and
190 ppm). Al and Fe decrease near the top of Unit A (A-1) (1038 and
1415 ppm) while values of Ca, Mg, Na, and K increase (312, 39, 66, and
226 ppm). The basal sediments of Unit B, as with Unit A, have lower
concentrations of Ca, Mg, Na, and K (198, 18, 54, and 244 ppm) than
sediments near the top of the unit (319, 28, 111, and 383 ppm). Fe in
Unit B decreases from 1842 ppm near the base to 1117 ppm near the
top although this is not mirrored by a decrease in Al, which increases
slightly from 1160 in the lower sediments to 1553 ppm in the upper. P
is noticeably higher (2433 ppm) in the basal sediments of Unit A
where the oldest ages for use of the shelter were obtained and may
be an indication of human activity. The concentration of P drops to
1225 ppm in the upper part of Unit A where no evidence of occupation was found. Levels of all elements are much higher in Unit C,
which has animal dung, abundant charcoal and a layer of ash as well
as mineral grains. The organic remnants certainly explain the high

values and suggest that high values lower in the sediment sequence
could also record human use of the rockshelter.
Charcoal concentrations from La Gruta 1 in Unit A have provided
the oldest evidence of human presence in the area and indicate an
earliest occupation between ca. 12,799e12,653 cal BP
(10,845  61 14C BP) with the rst period of occupation lasting from
about 12,799 to 12,049 cal BP (10,845  61e10,477  56 14C BP)
(Table 1; Franco et al., 2010). Charcoal concentrations revealed by
the excavation were limited in extent, being typically up to
10  20 cm in extent, although the deepest is partially covered by
the large boulder visible in the north wall of the excavation, which
has not yet been removed (Fig. 4). The charcoal scatters are also
separated vertically (i.e. in time) and spatially (i.e. in location)
despite the oor area of the shelter being small, although it may
have been more extensive in the past. These observations suggest
that the rst humans only used the rockshelter occasionally so that
occupation of the shelter was discontinuous over the 250e750 year
period indicated by the charcoal ages (12799e12549 cal BP 250
years; 12799e12049 cal BP 750 years).

Table 2
Radiocarbon ages for organic matter in lagoon sediments of La Gruta area.

Lagoon

Laboratory ID

La Barda

UGAMS 14755
UGAMS 14754
UGAMS 14759
UGAMS 14758
UGAMS 14700
UGAMS 14700

La Gruta 1
La Gruta 2

Depth
(cm)
10-15
35-40
14
55
25
65

Sample ID

13

EB-8
EB-3
GRUT2-2
GRUT2-1
LGLAG2-4
LGLAG2-1

-24.4
-24.9
-24.9
-24.9
-24.5
-25.0

Age
(14C BP)
2250 30
5690 35
2660 30
4900 30
2820 30
4710 35

Calibrated Age
(2 cal BP)
2329-2150
6525-6311
2842-2548
5657-5482
2958-2783
5576-5310

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

Radiocarbon ages for charcoal in sediment Unit A, at the base of


the excavation, date it to the PleistoceneeHolocene transition.
Small akes and charcoal concentrations were identied in the
deeper sediments of Unit A and at different locations within Unit B.
Artifacts of local and non-local material were recovered, the latter
including grey obsidian most likely transported from Pampa del
Asador (Franco et al., 2011). Early Holocene use of the rockshelter is
indicated by an age of 9029e8774 cal BP (8090  30 14C BP) on
charcoal from the base of a hearth near the top of Unit A (Site (f) in
Fig. 4). However, charcoal from the top of this hearth dated to
3487  38 14C BP (3832e3594 cal BP) showing that huntergatherers re-used the specic location of the old hearth later on
in time. The substantial difference in age between the lower and
upper sections of this hearth suggest either a hiatus in sedimentation or a period of erosion between the deposition of Units A and
B. It is possible that the upper section of the hearth is even younger
than the age we obtained because older charcoal may have been
incorporated into the new hearth when it was created at the same
location as the old hearth. Therefore, we consider the age of
3487  38 14C BP (Site (g) in Fig. 4) to be a maximum age for the
upper hearth, and also for the basal sediments of Unit B in which it
is located. A hearth near the top of Unit B (Site (k) in Fig. 4) has
provided an age of 400  20 14C BP (493e327 cal BP or AD 1457e
1626) for the youngest sediments in Unit B, indicating deposition
during the Little Ice Age (LIA) (Table 1).

probably a well-documented ood in 1983, deposited otsam into


the entrance area of the rockshelter and on nearby slopes (Fig. 11b).
The fact that the lagoon may be able to ood this rockshelter means
that ood deposits could be present in the sediments and also that
there could have been erosion of some deposits that may have
contained evidence of humans, during past oods. However, a line
of collapse ceiling blocks between the drip line of the shelter and
the back wall provide a protected area in front of the back wall
where sediments would be protected from erosion.
A test pit in La Gruta 3, begun in 2012, reached bedrock at
approximately 60 cm depth. It is close to the rear of the rockshelter
and behind large sandstone blocks that collapsed from the ceiling
(Fig. 11a). The sheltered location may partly explain preservation of
the sediment sequence in this part of the rockshelter. Six sediment
units (1e6) could be differentiated (Figs. 7e9). Units 1 and 2 have
similar texture containing 3.3% and 3.1% gravel and 33.2% and 32.7%
silt clay, respectively. They are both light yellowish brown slightly
gravelly muddy sand with a Folk and Ward (1957) mean grain size
in the very coarse silt range. Both deposits are very poorly sorted,
ne skewed and Unit 2 is mesokurtic and basal Unit 1 platykurtic.
The overlying Unit 3 is the nest of the six units with 4.7% gravel
and 39.5% silt clay. It is light yellowish brown slightly gravelly
muddy sand with a mean grain size in the very coarse silt range. It is
very poorly sorted, ne skewed, and platykurtic.
Elemental concentrations at LG3 differ from those at LG1, largely

Table 3
Color and texture of La Gruta 1 sediments.

Color
Unit

C
C
B
B
B
A
A

Sample
ID
LG1-6
LG1-5
LG1-4
LG1-3
LG1-8
LG1-1
LG1-2

LG1-7
Mean:
Sorting:
Skewness:

Kurtosis:

Verbal
Description

Munsell
Value

Gravel
%

Sand
%

Silt
%

Clay
%

Verbal
Description

Folk and Ward Statistics and Verbal Descriptions


Mean
(um)

Sorting

Skewness

Kurtosis

Verbal based
on Mean

muddy
sandy gravel
559.6
9.9 (vps)
-0.79 (vfs)
1.20 (l)
coarse sand
muddy
dark gray 10YR4/1
50.2 32.1 13.5
4.1 sandy gravel
645.1
9.0 (vps)
-0.81 (vfs)
1.1 (m)
coarse sand
dk yellowish
gravelly
brown 10YR4/4
26.0 41.3 19.6 13.1 muddy sand
145.8
15.4 (vps) -0.38 (vfs)
0.6 (vp)
fine sand
dk yellowish
gravelly
brown 10YR4/6
19.7 53.5 17.9
8.8 muddy sand
200.7
12.1 (vps) -0.43 (vfs)
0.8 (p)
fine sand
brownish
muddy
yellow 10YR6/6
39.7 37.9 14.8
7.6 sandy gravel
373.7
11.5 (vps) -0.68 (vfs)
0.9 (m)
medium sand
yellowish
muddy
brown 10YR5/6
40.0 32.8 17.5
9.7 sandy gravel
278.1
13.6 (vps) -0.64 (vfs)
0.7 (p)
medium sand
yellowish
muddy
brown 10YR5/8
35.4 39.8 15.4
9.5 sandy gravel
278.1
13.6 (vps) -0.64 (vfs)
0.7 (p)
medium sand
yellowish
muddy
brown 10YR5/6
32.3 39.1 20.0
8.5 sandy gravel
279.5
12.7 (vps) -0.59 (vfs)
0.8 (p)
medium sand
Graphic mean grain size is calculated as the size of the 16th, 50th, and 84th percentile divided by 3.
Very poorly sorted sediments are those in which grain sizes are mixed (large variance) whereas well sorted indicates that the sediment sizes are
similar (low variance). In Table 3 and 4 vps = very poorly sorted; ps = poorly sorted.
A grain size distribution may be symmetrical (s) or skewed. If the bulk of the data is at the left and the right tail is longer, the distribution is
positively or fine skewed indicating a tail of fine grains; if the peak is to the right and the left tail is longer, the distribution is negatively or
coarse skewed indicating a tail of coarse grains. In Tables 3 and 4 vfs = very fine skewed; fs = fine skewed; s = symmetrical.
In a distribution of grain size kurtosis is the height and sharpness of the peak relative to the rest of the data. Higher values indicate a higher,
sharper peak and lower values a lower, less distinct peak. A low, broad peak is platykurtic (p), a high, sharp peak is leptokurtic (l), and a
normal peak is mesokurtic (m). In Tables 3 and 4 vp = very platykurtic; p = platykurtic; m = mesokurtic; l = leptokurtic; vl = very
leptokurtic.
dark gray

10YR4/1

48.2

32.7

14.4

4.6

3.2. La Gruta 2 and 3


La Gruta 2 and 3 rockshelters have formed in Monte Len Formation sandstone along the margin of a complex lagoon (Figs. 2a
and 3b). The test pit at La Gruta 2 reached bedrock and provided
an oldest age for a guanaco bone showing human action of 8403e
8208 cal BP (7560  30 14C BP). Some 50 m away, the oor of La
Gruta 3 lies close to the level of the lagoon oor and a major ood,

because of the different substrates: fossilierous, carbonate-rich


sandstone and silicied ignimbrite, respectively (Figs. 6 and 9).
Levels of Al, Fe, and Mn are much lower at LG3 while Ca, Mg, Na, K,
Sr and Si are higher. Basal Units 1e3 in the Main Excavation at LG3
have generally higher Al, Fe, Mn and Si concentrations than overlying Units 4e6, while the opposite is the case for Ca, Mg, Na, K, and
Sr, with these elements increasing in the upper units. Elemental
levels in the Entrance Pit, near the drip line of the rockshelter, do

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

not differ signicantly from those in the lower units of the Main
site.
The basal Unit 1 sediments contained puma bones, including a
rib fragment, two thoracic vertebrae, and a second and a third
phalanx (ungual). Collagen in the rib bone provided an age of
36,934e36,136 cal BP (32,650  140 14C BP). Sediment samples
were collected at intervals through the prole for pollen analysis
but insufcient pollen was extracted from Unit 1 to allow useful
counts for paleoenvironmental reconstruction. Units 2 and 3 also
contained animal bones, namely of guanaco and Mylodontidae
(Figs. 7 and 10). In addition, a rib fragment of an unknown animal
was recovered as well as a bone that could not be identied.
Collagen from a guanaco phalanx from Unit 2 dated to 12,711e
12,562 cal BP (10,720  30 14C BP) and was stained by manganese,
suggesting exposure to water (Site (2) in Fig. 7).
The Mylodontidae bones included a rib, fragments of a thoracic
vertebra, and remains that could be from a pelvis or sacrum. A
fragment of a Mylodontidae thoracic vertebra (523/1) and two
other Mylodontidae bone fragments (523/2) from Unit 2 (Table 1;
Site (3) in Fig. 7) provided bioapatite ages of 9560  30 and

9470  30 14C BP (11,077e10,678 and 10,748e10,571 cal BP). These


ages are statistically identical at the 2s uncertainty level and
overlap when calibrated even at the 1s level. They also have similar
d13C values of 12.2& and 11.6& suggesting that the ages are
reliable. A bioapatite age was also obtained for a third Mylodontidae bone, a fragment of vertebra that was found in horizontal position (suggesting that it may not have been affected by
bioturbation) in sediment Unit 3 (516 in Table 1; Site 5 in Fig. 7).
This age of 8540  30 14C BP (9539e9466 cal BP) is younger than
those from Unit 2 and the bioapatite has a higher d13C value
of 9.5&.
Based on bone size, fusion state, and external morphology, the
Mylodontidae bones from Units 2 and 3 at LG3 appear to be from a
single individual. If correct, this suggests contamination of one or
more of the dated bone samples by older or younger carbon that
was not eliminated by laboratory pre-treatments for AMS radiocarbon dating. The two oldest ages are for different bone fragments
from the same part of Unit 2; the ages are identical (11,077e10,678
and 10,748e10,571 cal BP), so we consider them to be reliable. We
cannot entirely rule out that the Mylodontidae bone from Unit 3 is
from a different individual, and if so this might explain the younger
age from a bone in a younger sediment unit. If all the bones are
from only one individual, then this bone is the most likely to have
been contaminated. In any event, the three ages suggest that
Mylodontidae was in the area during the late Pleistocene to early
Holocene, a conclusion that is supported by the previously
mentioned late Pleistocene collagen age of 10,720  30 14C BP
(12,711e12,562 cal BP) for a guanaco phalanx bone that came from
sediment Unit 2 (Site (2) in Fig. 7).
Pollen data for Unit 2 indicate grass steppe with dwarf shrubs
(Ephedra, Nassauvia) so that conditions were probably slightly
wetter than today. To date, we have found no clear evidence of
human presence during deposition of Units 1 and 2. It is possible
that the Mylodontidae bones in Unit 3 were eroded from Unit 2 as a
result of post depositional processes, as they were found very close
to a pit. Supporting this is the evidence that the bones appear to be
from one individual. The unidentied rib fragment has lost part of
the cortical surface due to abrasion. This could have occurred as a
result of water action in the past possibly when the lagoon
extended into the shelter or when seepage waters entered the back
of the cave along ssures and bedding planes. There is a small
spring a few meters west of the shelter that was used by the

Fig. 8. Variations in sediment texture at La Gruta 3 rockshelter (top: Main Excavation;


bottom: Entrance Pit). For clarity, histograms for adjacent units are shaded differently.
See Fig. 7 for abbreviations and sample locations in the prole.

Fig. 9. Variations in elemental concentrations between units at La Gruta 3 rockshelter.


Values are in parts per million (ppm). Note the different scales in the upper and lower
diagrams.

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

previous owner of Estancia La Gruta for fresh water. After heavy


rain, similar seepages may develop in the rockshelter itself and
could have contributed to the uneven upper surfaces of some units
by eroding channels in them. A few isolated, small, worked akes
were recovered from Unit 3, but their origin has not yet been
determined.
Close to the drip line of La Gruta 3 rockshelter, at a lower
elevation than the Main Excavation site, there is a distinct lightcolored layer of sediment rich in gravel and sand that is cemented by secondary carbonate (Fig. 11a). Conditions at the time the
sediments accumulated may have been very like the lake conditions shown in Fig. 12a and b. A 50  50 cm pit (Entrance Pit; Fig. 7
and Table 4) was excavated in this deposit to establish stratigraphy
and to obtain material for dating.

Bones were recovered from 8.5, 16.5, 18, and 19 cm depth in the
sequence and bone fragments from 18 cm were dated. These provided a bioapatite age of 24,410  35 14C BP (28,654e28,211 cal BP).
The d13C value for the bioapatite is 8.9&, which appears high,
suggesting the possible incorporation of inorganic carbon from the
calcareous sandstone of the rockshelter. If this did occur, then the
estimated age of the bone should be considered a maximum value
as incorporated carbon would have virtually no 14C, making the
bone appear older than its true age. Using a tape and inclinometer,
we determined that the old surface of the cave oor at the back of
the rockshelter, where the Main Excavation was undertaken, is
0.735 m above the Entrance Pit deposit. As the Main Excavation
reached bedrock at w0.55 m this means that the upper surface of
the Entrance Pit deposit is only 18.5 cm below the base of the

Table 4
Color and texture of La Gruta 3 sediments (see Table 3 for list of abbreviations).

Color
Unit

Sample
ID

6b (surf
gravel)
6a

LG3-f2

LG3-e

4b
(sand)
4a

LG3-d2

LG3-c

2b (silt)

LG3-b2

2a

LG3-b1

LG3-a

EC

LG3E-3

EB

LG3E-2

Verbal
Description

Folk and Ward Statistics and Verbal Descriptions


Munsell
Value

Gravel
%

Sand
%

Silt
%

Clay
%

Verbal
Description

Mean
(um)

Sorting

Skewness

Kurtosis

multiple sand
colors
light yellowish
brown
very dark
gray**
light gray

N/A

12.0

87.0

0.9

0.0

gravelly sand

911.5

2.2 (ps)

-0.14 (fs)

1.20 (l)

2.5Y/6/4

9.9

75.5

7.2

7.3

289.7

5.7 (vps)

-0.11 (fs)

1.31 (l)

2.5Y/3/1

8.7

64.0

16.0

11.2

143.1

9.4 (vps)

-0.27 (fs)

0.88 (p)

2.5Y/7/2

0.2

91.0

5.0

3.7

116.5

2.2 (ps)

-0.21 (fs)

1.82 (vl)

light yellowish
brown*
light yellowish
brown,
light yellowish
brown
light yellowish
brown
light yellowish
brown
grayish brown

2.5Y/6/4

8.4

58.6

18.3

14.7

90.4

10.1 (vps)

-0.22 (fs)

0.79 (p)

2.5Y/6/4

4.7

55.7

20.1

19.4

62.3

8.8 (vps)

-0.20 (fs)

0.70 (p)

2.5Y/6/4

5.9

61.1

15.7

17.3

44.7

7.3 (vps)

-0.25 (fs)

1.06 (m)

2.5Y/6/4

3.1

64.1

14.5

18.4

54.00

7.6 (vps)

-0.28 (fs)

0.88 (p)

2.5Y/6/4

3.3

63.5

17.5

15.7

46.4

6.8 (vps)

-0.28 (fs)

1.02 (m)

2.5Y5/2

5.7

56.6

19.2

18.4

61.2

11.8 (vps)

-0.25 (fs)

0.64 (vp)

light brownish
gray
EA
LG3E-1
light yellowish
brown
*different colored gravels
** visible organic matter

2.5Y6/2

3.3

73.1

12.3

11.2

57.1

6.2 (vps)

-0.29 (fs)

2.53 (vl)

2.5Y6/4

9.7

63.7

15.5

11.2

gravelly muddy
sand
gravelly muddy
sand
slightly gravelly
sand
gravelly muddy
sand
slightly gravelly
muddy sand
gravelly muddy
sand
slightly gravelly
muddy sand
slightly gravelly
muddy sand
gravelly muddy
sand
slightly gravelly
muddy sand
gravelly muddy
sand

93.6

10.8 (vps)

-0.08 (s)

1.05 (m)

LG3-f1

LG3-d1

Excavation was difcult because the material was heavily


cemented in some layers and was only possible using a geologic
hammer. Weathered and fractured bedrock was reached at about
21 cm depth and 3 units EA, EB, and EC were exposed. At the base is
yellowish brown gravelly muddy sand of Unit EA, 6 cm thick,
containing 28.5% gravel and very coarse and coarse sand and 26.6%
silt and clay. Mean grain size is very ne sand, the unit is very
poorly sorted, and the particle size distribution is symmetrical and
mesokurtic. The overlying Units EB and EC (10 and 6 cm thick
respectively) are light brownish grey slightly gravelly muddy sand
and grayish brown gravelly muddy sand both with a mean grain
size of very coarse silt. These units contain 12.0% and 22.8% gravel
and very coarse and coarse sand, respectively. Units EB, and EC
contain 23.5% and 37.6% silt and clay, are very poorly sorted, ne
skewed, and very leptokurtic and very platykurtic, respectively.
Signicantly, Unit EB with 41.6% ne sand was the nest of the
three units and also the most cemented.

Verbal
based on
Mean
coarse sand
medium
sand
fine sand
very fine
sand
very fine
sand
very coarse
silt
very coarse
silt
very coarse
silt
very coarse
silt
very coarse
silt
very coarse
silt
very fine
sand

excavation where puma and Mylodontidae bones were found.


Therefore, it is possible that Entrance Pit Unit EA and Unit 1 at the
Main Excavation are of the same age. The possibility of this is
perhaps increased by sediment characteristics. If we assume that
the medium sand to clay fractions of each deposit total 100%, then
Unit 1 has 62.6% medium, ne and very ne sand, and 37.4% silt and
clay. In comparison, Unit EA has 62.7% medium, ne and very ne
sand, and 37.3% silt and clay, virtually identical to Unit 1. In terms of
the ner grain components therefore, the deposits are identical and
so could represent one unit, which extended from the front to the
back of the rockshelter.
The basal Unit 1 sediments containing the puma bones (36,934e
36,136 cal BP) are separated considerably in age from the overlying
Unit 2 sediments with Mylodontidae and guanaco bones dating in
the range 12,711e10,571 cal BP (10,720  30e9470  30 14C BP)
(Table 1). The upper surface of Unit 1 in the southern wall of the
excavation, near the back wall of the rockshelter, shows very little

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

10

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

relief (<5 cm). In contrast, the upper surface of Unit 2 has a relief of
around 10 cm suggesting some erosion of the sediments after
deposition. This would support the idea proposed earlier, that the
Mylodontidae bones in Unit 3 were eroded from Unit 2.
4. Evidence of past water levels in La Gruta area lagoons
Flood debris near La Gruta 3 rockshelter provides evidence of
past ooding events in the adjacent lagoon. However, to obtain
additional information on past conditions in lagoons in La Gruta
area, whether they contained water at particular times and how
high lake levels were, we examined sediments in La Gruta Lagoons
1 and 2 and in a nearby lagoon La Barda (Fig. 2). La Gruta 1 rockshelter formed in a steep cliff along the southern margin of La Gruta
Lagoon 2 while La Gruta 2 and 3 rockshelters formed in steep cliffs
bordering the southern margin of La Gruta Lagoon 1.
La Gruta Lagoon 2 is about 900 m long from west-to-east and
consists of three separate basins separated by rock-cored ridges
mantled by beach gravels that preserve relict shorelines of former
higher lagoon levels. The largest basin is to the west and it is 470 m
long and up to 200 m wide. To the east is a second basin 125 m WeE
and 170 m wide, and east of this a third that is 150 m WeE and up to
90 m wide. Small streams up to 1 km long enter from the west and
south but these have very small catchments. When full, the large
basin to the west may have been 3e4 m deep.
La Gruta Lagoon 1 is 1.4 km NNWeSSE and a maximum of
1.05 km WeE. The lagoon depression has two separate basins that
ll independently but as water levels increase they coalesce to form
a single lake. The northern depression is the deepest and is 775 m
NeS and 750 m WeE. The second basin in the southwest is 536 m
NeS and 362 m WeE. Relict beach ridges and shorelines around the
depression but mainly on the east and southeast margins (the
downwind side) show that when full the lake would be up to 10 m
deep in the northern depression and 6e7 m deep in the southern
depression. In the broad at north of La Gruta 3 rockshelter the
water would be 6 m or more deep. La Gruta Lagoon 1 differs from
Lagoon 2 in that it is fed by a stream that extends to the north and
west for about 18 km. The channel of this stream enters the eastern
side of the lagoon depression and winds across the oor to the
deeper northern depression. Recent extensive ooding of La Gruta
Lagoon 1 is presumably a result of this large stream input and this
must also have occurred in the past and been more frequent during
periods of increased moisture.
La Barda is a 375 m long, triangle-shaped depression elongated
WeE; it is broad in the west and narrows to a point in the east.
There are two distinct basins separated by an arcuate 1.5 m high
gravel beach bar produced by wave action generated by the prevailing westerly winds when the large basin was ooded. The basin
to the west is 225 m WeE and 225 m across at its widest point; the
smaller basin to the east is 150 m long and at its widest only 50 m
across. When ooded, the westerly basin is probably occupied by a
lake 2e3 m deep.
In most years, all three lagoons dry out during the dry season,
allowing examination of sediments in the basins. However, based
on satellite evidence water persists in La Gruta Lagoon 1 longer
than Lagoon 2, which in turn contains water longer than La Barda.
During 2012 and 2013, excavations in the sediments covering the
oors of these three lagoons provided information on their age and
therefore on when the lagoons contained water. At La Barda (48
49.7960 S; 69 31.9900 W; 319  5.5 m elevation) an excavation in
the middle of the larger basin to the west exposed sediments to
50 cm depth with the basal material being a coarse gravel. Samples
were taken continuously from 5 cm increments making 10 samples.
The samples from 35 to 40 cm and 10e15 cm below the surface
provided AMS radiocarbon ages of 5690  35 and 2250  30 14C BP

(6525e6311 and 2329e2150 cal BP). A pit was also excavated at the
approximate center of La Gruta Lagoon 2 (48 49.4730 S; 69 23.9390
W; 275  4.9 m elevation) to a depth of 65 cm. Samples from the
base of this excavation and from 25 cm depth provided AMS
radiocarbon ages of 4710  35 and 2820  30 14C BP (5576e5310
and 2958e2783 cal BP). A third pit 60 cm deep was excavated in a
relict beach/river bar to the north of La Gruta 3 rockshelter (48
50.2340 S; 69 22.3770 W; 259  4.9 m elevation). Organic material
from 55 cm and 14 cm depth provided AMS radiocarbon ages of
4900  30 and 2660  30 14C BP (5657e55482 and 2842e
2548 cal BP).
Excavations in all three lagoons reached gravel suggesting that
prior to sediment deposition, which began around 6500 cal BP,
there was a period of dry, windy conditions with deation of ne
sediments from the lagoon basins. Ages of 2329e2150, 2958e
2783 and 2842e2548 cal BP (2250  30, 2820  30, and
2660  30 14C BP) from 10 to 15, 25, and 14 cm depth suggest
deposition of ne sediments in water up to at least 2100 cal BP.
The implication of these results is that conditions were much
drier than today prior to 6500 cal BP and at least as wet as today
after this date. This suggests that after humans rst entered La
Gruta area around the time of the PleistoceneeHolocene transition, there was a major dry interval of climate. The timing of this
dry interval corresponds closely with the lack of evidence for
human occupation of the three La Gruta rockshelters around this
time. Therefore, humans may have entered this area when water
was reasonably abundant in the lagoons. They may then have
used the area less frequently in the Holocene between ca. 8000e
6500 cal BP due to much drier conditions when the lagoons were
largely bare windblown surfaces with no water, as indicated by
the basal gravels in all three excavations. Human groups may have
moved to nearby areas with more reliable water supplies during
this period.
In 1983, there was signicant ooding of La Gruta Lagoon 1 to
the extent that the lagoon inundated a farmhouse on Estancia La
Gruta just west of La Gruta 3 rockshelter (Fig. 12a and b). The
house is at about the same elevation as the oor near the entrance
to La Gruta 3 rockshelter. Driftwood associated with the ood is
widely distributed around the east and southeast margins of the
lagoon basin, striking evidence of how high the waters reached
(Fig. 11b). Historical photographs show the extent of the ooding
which caused the farmer to re-locate the house and associated
structures; these photographs are an indication of what the area
may have been like in the past when the climate was somewhat
wetter than today or when there were heavy rainfall or snowmelt
events. There appears to have been above-average snowfall prior
to the 1983 ood (Florence Kemp, Estancia 17 de Marzo, pers.
comm. 2014), and melting snow may have been a large component of the oodwaters.
The images of the ood in Fig. 12a and b clearly show water
carrying a great deal of sediment in suspension. As oodwaters
subside, any suspended silt and clay accumulates on the oor of the
lagoon and in sediments around the shore (including in the rockshelter if waters are high enough) when there is limited wave action. If the prevailing southwesterly winds strengthened when the
lagoon was ooded, beach bars would form due to wave action on
the shore in the downwind direction. We would therefore expect
that an increase in nes in the rockshelter sediments might record
a period of increased moisture when ooding was more frequent
and the extent and duration of oods greater than today. The 1983
ood also shows what conditions may have been like in the past,
and that occupation of the shelter would have been difcult when
ooding was more frequent and more prolonged. The level of water
in 1983 also shows that sediments in the back of the rockshelter
could have been eroded by waves, particularly during fall as the

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

11

Fig. 10. Mylodontidae bones from the Main Excavation at La Gruta 3 rockshelter.

southwesterly wind belt migrated northwards and strengthened as


winter approached.
5. Interpretation of La Gruta rockshelter sediment sequences
The sediments in a rockshelter commonly result from simultaneous inputs from geogenic and biogenic sediments due to a variety
of processes such as rockfall, wind action, and human activity
(Farrand, 2001). Geogenic sediments may originate either inside
the cave (endogenic) or outside (exogenic) and can include: roof fall
(either spalling or collapse); disintegration of the cave bedrock by
chemical and freeze-thaw weathering; inwashing of colluvium and

soil through the cave mouth or through ssures leading to the cave;
fallout of aeolian sand- or silt-sized dust; stream deposition, from
waters entering the cave via ssures or bedding planes or from
external streams owing into the cave mouth; beach sand or gravel,
if the cave is now or was close to a lake. Granulometric histograms
showing the particle-size distribution of the sediment can reveal
modes in particle size that can help to establish the dominant
processes responsible for sediment being transported to the rockshelter. A very coarse mode is likely to reect breakdown of the
cave ceiling possibly by freeze-thaw weathering, a silt mode
probably aeolian material blown into the cave, and a clay mode
input of soil through cracks and ssures in the rockshelter walls and

Fig. 11. Evidence of ooding near La Gruta 3 rockshelter. Cemented muddy sandy gravel at the entrance to the shelter being excavated (Entrance Pit). The Main Excavation is below
the seated gure at the back of the shelter (a). Branches deposited along the shore of La Gruta Lagoon 1 to the east of La Gruta 3 rockshelter most likely during the 1983 ood (see
also Fig. 12). La Gruta 3 rockshelter is just beyond the point of the sandstone cliff in the upper left of the photograph (b).

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

12

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

Fig. 12. Flooding of La Gruta Lagoon 1 in 1983 and how this affected La Gruta 3 rockshelter. It is clear that the rockshelter could have been ooded at this time. Photographs (a) and
(b) were provided by Carlos Baetti who obtained them from the previous owners of Estancia La Gruta. We use them here with Baettis permission.

ceiling or a period of surface stability and soil development in the


cave (Farrand, 2001). A diffuse sand mode can indicate introduction
of beach deposits into the shelter. However, a coarse mode of
angular rock fragments may also be produced by hydration
weathering or the frequent wetting and drying of a rock surface.
Periods of relative stability and weathering of rockshelter sediments can often be interpreted from color and clay content. Dark
brown or reddish brown (oxidized) colors in a sediment sequence
that is generally light-colored can indicate a weathered horizon,
while enrichment in clay content relative to overlying and underlying strata is also evidence of pedogenic alteration of primary
minerals into clay minerals. Finally, bioturbation by plant roots and
burrowing animals, as well as pits dug by humans for the
emplacement of hearths or burials, obscure the original stratication. Farrand (2001) notes that the occupants of the Abri Pataud
rockshelter (Dordogne, France) removed sediment from under the
overhanging rockshelter where they were living, creating a
depression along the back wall at least 75 cm deep that truncated
older strata (see also Farrand, 1975a,b).
The granulometric histograms for La Gruta 1 and 3 rockshelters
reveal multiple modes in size frequency (Figs. 5 and 8) as do the
data in Table 5 of summary statistics for each of the three sediment
sequences examined (LG3 Main, LG3 Entrance Pit, and LG1). Table 5

reveals that LG1 sediments (gravel to clay size) have modes in the
gravel (36.5%) and silt (16.6%) size ranges. This differs from the LG3
sequences where the peaks are in the coarse sand (12.1% Main and
7.65% Entrance) and very ne sand (17.75% and 27.2%) ranges. If we
consider only particle sizes ner than very coarse sand, and convert
these to 100%, then the modes at LG1 are coarse sand (17.8%) and
silt (32.5%) and at LG3 coarse sand (14.6% Main and 8.8% Entrance)
and very ne sand (21.3% and 31.4%). La Gruta 1 rockshelter is a few
meters above the oor of the adjacent lagoon and is never ooded
as it is higher than the overow for the lagoon. As a result, it is
dominated by gravel produced by breakdown of the walls and
ceiling of the shelter and by silt that is blown into the shelter by
wind. In contrast, La Gruta 3 is close to the level of the adjacent
lagoon basin oor and is known to ood after heavy rains, as it did
in 1983 (Fig. 12a and b) so that sediments there have modes in the
coarse and very ne sand ranges although they still contain high
proportions of silt and clay. The elevations of relict beach ridges
around the lagoon suggest that it may have ooded beyond the
1983 level in the past, in which case it is possible that the oodwaters would have inundated parts of La Gruta 3 rockshelter. If the
shelter was ooded occasionally, then sediments in its oor may
preserve evidence. The dominance of sand at La Gruta 3 is also
related to the sandstone bedrock, which breaks down to produce

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

13

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

sand-sized particles more readily than the silicied ignimbrites at


La Gruta 1.

Unit B, similar to the semiarid conditions of today (Mancini et al.,


2013; note that in this paper LG1 units are numbered not dened

Table 5
Textural frequency modes in the sediment sequences at La Gruta 1 and La Gruta 3 rockshelters. Modes are indicated by shading.

Sediment Size
Category
Gravel
Very coarse sand
Coarse sand
Mdium sand
Fine sand
Very fine sand
Silt
Clay
TOTAL

La Gruta 3
Main Excavation (%)
Entrance Pit
A*
B*
A*
7.0
6.3
10.0
7.2
12.1
14.6
7.6
10.1
12.2
7.2
16.2
19.5
15.3
17.7
21.3
27.2
13.8
16.6
15.7
13.1
15.8
13.6
100%
100%
100%

(%)
B*

8.8
8.3
17.7
31.4
18.1
15.7
100%

La Gruta 1 (%)
A*
36.5
12.4
9.1
6.7
5.2
5.2
16.6
8.2
100%

B*

17.8
13.1
10.2
10.2
32.5
16.1
100%

* A: 100% includes gravel and very coarse sand; B: 100% excludes gravel and very coarse sand.

Unit A, at the base of the LG1 sediment sequence, has a darker


color (yellowish brown) than other units. It has high gravel and clay
contents, the former suggesting that water was available for hydration breakdown of the shelter walls or for freeze-thaw weathering. The color and high clay content suggest stable conditions in
the cave and pedogenesis due to increased moisture availability.
Unit B contains less gravel than Unit A and as a result sand, silt and
clay percentages are reasonably high, particularly silt. Overall, we
believe that conditions were somewhat drier during much of the
period when Unit B was deposited. The lower sediments of Units A
and B have higher Al and Fe and lower Ca, Mg, Na and K than the
sediments above them, suggesting wetter conditions when they
accumulated leading to some leaching of highly soluble elements.
Lower levels of Al and Fe and higher levels of Ca, Mg, Na, and K in
the top sediments of these two units indicate drier conditions and
possibly a stable surface in the rockshelter for some considerable
time. Pollen assemblages in Unit A (samples 1e8 in Fig. 4) and in
the basal sediments of Unit B (samples 9e10) have high percentages of grass pollen attesting to wetter conditions. However, pollen
samples 3, 11 and 12 from the upper deposits of Units A and B have
lower levels of grass pollen, which together with higher levels of
soluble elements is a strong indication of drier conditions (Fig. 13).
The pollen data for Unit B show a trend to reduced grass and higher
shrub percentages towards the top of the unit.
The uneven upper surface of Unit A (lower around the large
boulder exposed by the excavation), its variable thickness, high
levels of soluble elements towards the top, and the apparent large
age difference between it and Unit B, all indicate a lengthy period of
surface stability with some erosion of the rockshelter oor, possibly
under wetter conditions than today. High grass pollen percentages,
higher Al and Fe and lower levels of easily-dissolved salts in the
lower parts of Unit B, suggest wetter conditions that may have been
partly responsible for erosion of the upper sediments of Unit A.
High levels of Ca, Mg, Na and K in the upper part of Unit B may
record a period of stability and a hiatus in sedimentation before
deposition of Unit C.
Unit C has the highest percentage of gravel in the sequence but
lower silt and clay levels than the other units, suggesting relatively
wet conditions. However, low percentages of medium and ne sand
point to a drier climate than during deposition of Units A and B
because the larger clasts were not broken down. Only one pollen
sample (13) was taken from Unit C and this was near the base of the
unit (Fig. 4). This records a drier climate than in the upper part of

by letters). The higher percentage of Chenopodiaceae in sample 13,


compared with samples 11 and 12 from the upper part of Unit B,
indicate lagoon desiccation, because these plants grow around the
margins of water bodies that undergo frequent drying. The increase
in Poaceae and reduction in the Chenopodiaceae in the surface
pollen sample suggest slightly wetter conditions towards the present with more frequent ooding of the lagoon (Fig. 13).
The sediments at La Gruta 3 differ from those at La Gruta 1 in
having modes in different size categories. This is due to differences
in bedrock at the two sites (fossiliferous sandstone at LG3 and
silicied ignimbrite at LG1) and also to the elevations of the rockshelters above the adjacent lagoons. LG1 is a few meters above its
lagoon, so that wind-blown, saltating sand particles near the
ground surface cannot reach the shelter and so aeolian deposits
here are largely of silt size. In contrast, because the oor of LG3 is
close to the level of the adjacent lagoon, saltating sand grains can
enter the shelter so that aeolian sediments here are largely in the
very ne sand range (Table 5). At LG1, there is a silt particle size
peak throughout the sediment sequence. In contrast, at LG3, ne
sand is only a particle size peak in Units 1 and 2 but not in the
overlying units. This suggests that while silt has been available for
aeolian transport throughout the late Pleistocene and Holocene (as
evidenced by the LG1 record), ne sand has not. We believe that
aeolian input of ne sand to LG3 only occurred when ne sand was
available in the basin of the adjacent lagoon, and because this
would have required increased chemical weathering of the sandstone bedrock and/or transport of sand by streams entering the
depression, more water would have been needed. Any increase in
stream ow to the LG3 lagoon could have been accompanied by
ooding of the depression, and this may have deposited ne sediment on the oor of the LG3 rockshelter.
Signicantly, Unit EA of the LG3 Entrance Pit sediment sequence
has similar characteristics to those of Unit 1 at the Main Excavation
and is broadly similar in age. Unit EA does have a slightly higher
percentage of coarser grains but this is to be expected given its
location close to the drip line of the rockshelter. Puma bones in Unit
1 date to ca. 36,500 cal BP and unidentiable bones in Unit EA to ca.
28,300 cal BP, placing deposition of these units to roughly the same
time period, namely before or during the Last Glacial Maximum
(LGM), which would have brought very cold conditions to the area.
Unit EB is also broadly similar in grain-size characteristics to Unit 2
at the Main Excavation and Unit EC is similar to Unit 3 (and also to
Unit 4). Units 2 and EB are both dominated by a marked mode of

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

14

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

Fig. 13. Stratigraphic units and pollen assemblages in LG1 and LG3 sediments. Pollen sample numbers are indicated and radiocarbon ages in

very ne sand and low frequencies of coarser grains. In contrast,


Units 3 and 4, like Unit C, lack a distinct mode of very ne sand but
have less distinct modes of silt and ne sand and contain larger
numbers of coarser grains ranging in size from medium sand to
gravel.
However, despite their location below the drip line of LG3, Units
EA and EB contain relatively few coarse grains indicating little

14

C BP are shown at left.

breakdown of the ceiling of the shelter and of the rocks on the


surface above it during deposition of these units. This implies that
conditions were no wetter than semiarid when Units 1 and 2, and
Units EA and EB, were deposited. Pollen sample 1 did not provide
enough pollen for realistic paleoenvironmental inferences to be
made about Unit 1. Absence of pollen could mean drier conditions
with a reduced vegetation cover but more work will be necessary to

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

test this assumption. The absence of coarser rock fragments could


also indicate much colder temperatures throughout the year.
Temperatures consistently below freezing would limit freeze-thaw
cycles, and therefore rock breakdown, even if water was sufcient.
The substantial age difference between Units 1 and 2, of about
24,000 years, suggests that the upper part of Unit 1 may have been
eroded before deposition of Unit 2.
LG3 Unit 2, with ages between 12,711 and 10,571 cal BP
(10,720  30 and 9470  30 14C BP), is late Pleistocene in age and
may be contemporaneous with Unit A at LG1, which has provided
ages between 12,799 and 8774 cal BP (10,845  61e
8090  30 14C BP). An increase in moisture would increase mechanical breakdown of the walls of the rockshelters at La Gruta,
producing coarse particles up to gravel size. Gravel is abundant in
Unit A at LG1, but larger rock fragments are less common in Units 1
and 2 at LG3, although they are more abundant in what we have
argued to be equivalent Units EA and EB, near the drip line of the
rockshelter, where more water would have been available. In many
rockshelters, the rear of the shelter is drier than areas nearer the
entrance and this could help to explain the absence of coarser
particles in Units 1 and 2. Pollen data for LG1 Unit A and LG3 Unit 2
indicate a predominance of grass with dwarf shrubs (Ephedra,
Nassauvia) (Fig. 13), which supports our interpretation of the
sediment data as indicating semiarid conditions but with more
available water than today.
The character of the sediments at LG3 changes with deposition
of Unit 3 and later Unit 4. While maintaining a predominance of
ner particles, although at reduced levels, these units have a more
even distribution of particle sizes compared to Units 1 and 2. This
trend towards increased percentages of coarser particles suggests a
transition to slightly drier conditions in the rockshelter with less
sand being produced by chemical weathering and/or transported to
the lagoon by streams and then blown into the rockshelter by wind.
However, the moisture level during deposition of Units 3 and 4 was
high enough to break down the rockshelter walls and ceiling,
contributing fragments of medium sand to gravel-size to the oor
of LG3. The trend to higher percentages of coarser fragments
continued through Units 5 and 6 with the particle size mode in Unit
6 being coarse to very coarse sand. The increase in the percentage of
coarser particles upwards in the sediment sequence was accompanied by lower concentrations of Al, Fe and Si, and higher concentrations of soluble elements Ca, Mg, Na, and K, both changes
suggesting drier conditions. Pollen data for Unit 3 at LG3 suggest
semiarid conditions similar to those of today. Current age data for
Unit 2 and for the upper part of Unit 3 place deposition of Unit 3
sediments between ca. 10,748 and 503 cal BP. If the age for the
Mylodontidae vertebra from Unit 3 is reliable, deposition may have
occurred between 9539 and 503 cal BP. At present, we do not know
if sedimentation was continuous, or if the age of 503 cal BP is for
charcoal resting on a much older surface. Therefore, a reliable
reconstruction of vegetation and environmental conditions for the
area during the Middle and Late Holocene will clearly require
further research.
The south wall of the excavation near the back wall of La Gruta 3
exposes a well-dened animal burrow or human pit lled with Unit
4 sediments that extended through Unit 3 into Unit 2 (Fig. 7). The
burrow or pit is close to where Mylodontidae bones were found in
the upper sediments of Unit 3 and human or animal disturbance of
the sediments may explain why Mylodontidae bones have been
found in both Units 2 and 3. The lling of the burrow or pit in Units
3 and 2 by Unit 4 sediments suggests that Unit 4, and probably also
Units 5 and 6, were deposited under relatively dry conditions when
humans and small animals could occupy the shelter.
Units 5 and 6 differ from Unit 4 in having larger percentages of
coarse grains and a more even distribution of grain sizes. In Unit 5,

15

sediment sample e1 has modes of coarse sand and silt while sample
e2 has a single broad mode of ne sand. Nearer to the back wall of
the cave Unit 6 contains more gravel (mode of very coarse to coarse
sand) presumably from breakdown of the low ceiling at this location. Based on four radiocarbon ages, Units 4e6 appear to have
accumulated during the last ca. 550 years. Pollen sample 10 from
Unit 5 has high shrub percentages (Asteraceae Asteroideae) and a
low proportion of grass pollen that suggest dry conditions. Pollen
sample 12 from Unit 6, with a higher percentage of grass pollen,
suggests a slight increase in moisture in recent times. The LG3 data
for the last 500 years complement that from LG1. Both records
reveal a semiarid environment similar to the present but with evidence of an early dry period followed by somewhat wetter conditions towards the present.
All of the sediment units in the Main Excavation of LG3 rest
against large collapse blocks but to the north, towards the rockshelter entrance, these sediments are buried beneath other collapse
blocks that produced the pile of boulders visible in Fig. 11a. The
upper surfaces of Units 2 and 3 are also very irregular suggesting
erosion by water and by animals prior to burial by the overlying
units. Bioturbation is particularly apparent in Unit 3 and in the
upper part of Unit 2, including channels and animal burrows or
human pits lled with Unit 4 sediments. Clearly, the back of the
shelter was not ooded when these activities occurred.
6. Discussion
Evidence from La Gruta area and from caves 25 km to the north
and northwest of the area, including La Martita Cave 4 (Aguerre,
2003), Viuda Quenzana 8 (Franco et al., 2013) and El Verano Cave
1 (Durn et al., 2003) (Fig. 3) indicates human presence at La Gruta
1, in the southern Deseado Massif, during the PleistoceneeHolocene transition at ca. 12,799e12,049 cal BP (10,845  61e
10,477  56 14C BP) and at other sites during the early Holocene at
ca. 9029e7760 cal BP (8090  30e7500  250 14C BP). At the time
of the initial human occupation of the area, vegetation near La
Gruta 1 resembled patches of grass that are found in the presentday dwarf-shrub steppe of the Deseado Massif above 700 m, suggesting cold and semi-arid conditions (Mancini et al., 2013) and
similar pollen evidence has been found in Unit 2 at La Gruta 3.
Mylodontidae bones have been discovered previously at Piedra
Museo approximately 55 km to the north of La Gruta, in sediment
units dating between 11,000  65 14C BP (12,994e12,713 cal BP)
and 9230  105 14C BP (10,655e10,180 cal BP) or 15,686e15,076 to
10,655e10,180 cal BP if the oldest date of 12,890  90 14C BP from
the site is accepted (Miotti et al., 1999; Miotti and Salemme, 2003).
The nding of Mylodontidae bones at La Gruta 3 extends the known
distribution of this giant ground sloth to the southern Deseado
Massif, and the ages we have obtained correspond with the age
range postulated for the Piedra Museo deposits. At Piedra Museo,
Mylodontidae coexisted with other extinct mammals and guanaco
and its presence in the archaeological record is seen as the result of
different hunting events (Miotti and Salemme, 2003). To the south
of La Gruta the closest site with Mylodontidae bones is a natural,
undated deposit in the Yaten Guajen canyon 140 km to the south
(Franco et al. 2007). South of Patagonia, extinct fauna have been
recovered from a number of sites, mainly in ltima Esperanza
Province in Chile but also closer to the Atlantic Ocean.
Mylodontidae darwini was herbivorous and has been linked
traditionally to open areas and a temperate to cold semiarid climate
(Brandoni et al., 2010). It is one of the few extinct ground sloths for
which preserved dung has provided direct evidence of the plants
eaten (Moore, 1978). Recent studies based on biomechanics and
functional morphology (Bargo et al., 2006a, 2006b; Bargo and
Vizcano, 2008) indicate that M. darwini was a mixed or selective-

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

16

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

feeder capable of selecting specic plants or parts of plants. The


dominant plants in Mylodontidae darwini dung in a cave at Ultima
Esperanza, in southernmost Patagonia, were grasses and sedges
(Moore, 1978), which supports the morphological information obtained. In her study of pollen and cuticles in nine dung samples
from Mylodontidae Cave in southern Chilean Patagonia, Markgraf
found that The pollen data revealed much greater plant diversity
than the cuticle data (Markgraf, 1985: 112). Heusser et al. (1992)
analyzed ve samples with good stratigraphic provenience dated
between 11,330 and 12,570 14C BP obtained by Borrero and recognized that the samples offered a mixture of ingested and windblown pollen and spores, but observed that the results were
concordant with the existence of a tundra or dwarf shrub heath at
the end of the Pleistocene. The paleoecological implication is that
the region was basically treeless at the time of the sloths. That was
also Nordenskjlds conclusion on the basis of the absence of leaves
in dung (Nordenskjld, 1996 [1900]: 111). The conclusion from
these botanical analyses is that open environments existed at the
end of the Pleistocene in Ultima Esperanza (see also Villa-Martnez
and Moreno, 2007), in an area where Mylodontidae remains were
recovered. Taking into account the plant assemblage from the El
Palmar Formation in northeastern Argentina and the feeding specializations considered for M. darwini, this species would have had
a broad range of suitable vegetation to select as food, from cold and
arid to warm and humid environments (Brandoni et al., 2010). At
several sites, possible evidence of human action on Mylodontidae
bones (e.g. cut marks) has been difcult to conrm (see for example
Borrero and Martin, 2008 for Las Buitreras Cave), being in some
cases related to animals (Martin et al., 2013; Prevosti and Martin,
2013).
The Mylodontidae bones at La Gruta 3 are younger than the
oldest ages for the presence of humans at La Gruta 1 only about 2 km
away. This suggests that humans and Mylodontidae were probably
both in the southern Deseado Massif during the late Pleistocene but
does not reveal whether they encountered one another. What is
apparent is that Mylodontidae continued to use the area even after
the rst humans had explored it. How humans and Mylodontidae
co-existed and why Mylodontidae eventually became extinct have
been subjects of intense debate (e.g. Borrero et al., 1997; Borrero,
2001; Long et al., 1998; Barnosky and Lindsey, 2010 and references
therein). To date, there is no evidence of human occupation of La
Gruta area between ca. 12,049 and 9029 cal BP (10,477  56 and

8090  30 14C BP), while our ages for the Mylodontidae bones date in
the range 11,077e9466 cal BP (9560  30e8540  30 14C BP).
However, there is evidence of occupation at La Martita Cave 4 and at
El Verano Cave 1, 25 km to the northeast (Aguerre, 2003; Durn et al.,
2003). This suggests that Mylodontidae was still in the area after
humans explored La Gruta and presumably moved elsewhere,
possibly because of drier conditions and a lack of a permanent
source of water. We have found no clear evidence of humans in the
area when Mylodontidae was there, or of human actions on Mylodontidae animals or bones. The absence of such information could
be related to discontinuous human presence in the area, and/or to
the small size of the human population at this time, as suggested by
other researchers (e.g. Borrero, 1994-95).
To date, only small akes associated with the nal stages of
stone tool manufacture have been found in the oldest deposits at La
Gruta 1. There is evidence for transport of grey obsidian, probably
from Pampa del Asador in southern Patagonia, about 175 km
northwest of La Gruta (Stern, 2000), or from its secondary area,
which extends 30 km to the east (Belardi et al., 2006). Translucent
chalcedony was also transported to LG1, probably from the Viuda
Quenzana area about 25 km to the north. However, surface artifacts
have been found that can probably be related to early human inhabitants due to their morpho-technological characteristics
(Fig. 14). The preform shown in Fig. 14a has characteristics that
resemble artifacts from Tres Arroyos, Tierra del Fuego, with radiocarbon dates of 10,280  110 and 10,420  100 14C BP (12,555e
11,407 cal BP) (Jackson, 2002). In addition, the preform of the stem
shown in Fig. 14b resembles Fell 1 projectile points which also have
old dates (e.g. Bird, 1988, Massone and Prieto, 2004).
The archaeological sites of La Gruta 1, La Gruta 2, La Martita Cave
4, and El Verano Cave 1 have produced evidence of human occupation in the period 10,294e7760 cal BP (8960  140 and
7500  250 14C BP) suggesting that conditions were wet enough to
sustain hunteregatherer activities at this time. During the early
Holocene, pollen evidence shows that grass was even more dominant in the steppe vegetation at La Gruta than during the late
Pleistocene suggesting even wetter conditions (e.g. Mancini et al.,
2013). However, after ca. 7760 cal BP (7500  250 14C BP) there is
no evidence of human occupation until 4770  25 14C BP (5583e
5325 cal BP) at Viuda Quenzana, at 4475  95 14C BP (5311e
4846 cal BP) at La Martita Cave 4 (Aguerre, 1982), and at
3487  38 14C BP (3832e3594 cal BP) at La Gruta 1 (Table 1; Franco

Fig. 14. Projectile point preform found on the ground surface at La Gruta (a) and stem of a projectile point preform found on the ground surface close to La Gruta area (b).

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

17

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

et al., 2013). The possible absence of humans in the southern


Deseado Massif from about 7760 to 5583 cal BP (7500  250e
4770  25 14C BP) agrees well with the sedimentological evidence
from the LG1, LG3 and La Barda lagoons, which indicates dry conditions prior to about 6500 cal BP (5690  35 14C BP) but somewhat

13,000 cal BP and slightly wetter conditions with grass steppe and
dwarf shrub vegetation from ca. 13,000 to 11,500 cal BP. The
vegetation changed to grass steppe after ca. 11,500 cal BP until ca.
8800 cal BP indicating wetter conditions during this period (Brook
et al., 2013; Mancini et al., 2013).

Table 6
Summary data on human occupation of La Gruta area. Ages in cal BP or cal AD (where indicated) Kilian and Lamy (2012).

SEDIMENTS1
Age (cal BP) or
Stratigraphic Unit
Units 4&5, LG 3

Charcoal/guanaco bones (h)

Unit C, LG 1

Charcoal

HUMANS3
Age (cal BP)

Correlations4

Period (cal BP)

539-156
(AD 1411-1794)
493-327
(AD 1457-1623)

Little Ice Age

539-156
(AD 1411-1794)

Medieval Climate Anomaly

1266-539
(AD 571-1411)

Higher temperatures Antarctic EDC ice core;


weaker Westerly winds.

6500-1270

Comments2

Dry interval no human occupation?


6525-2150
Unit B, LG 1
Unit B, LG 1

La Gruta Lagoons 1 & 2, La Barda


Charcoal
Charcoal
Charcoal at La Martita
guanaco bones (h) Viuda Quenzana

1881-1271
3832-3594
5311-4846
5580-5320

Hiatus in sediment deposition: LG1 Unit A to Unit B and LG3 Unit2 to Unit 3.
Pollen indicate drier conditions
guanaco bone (h) LG2
8403-8208
Charcoal
Unit A, LG 1
9029-8774
Charcoal La Martita
9429-8207
Charcoal El Verano
8972-7760
Charcoal El Verano
10,294-9551
Unit 2, LG 3; Mylodontidae/guanaco
12,700-10,600
bones
Charcoal
Unit A, LG 1
12,799-12,049

36,500

Entrance Pit Unit EA unidentified


bones
Unit 1, LG 3; puma bones

YD warming followed by early Holocene


temperature maximum in Antarctic EDC ice
core; weaker Westerly winds. Pollen data
indicate wetter conditions.

LGM including H1, H2, H3, ACR (1500013,000 cal BP); strong Westerlies. EPICA
Dome C major dust period.

Hiatus in sediment deposition: LG3 Unit1 to Unit 2.


Pollen indicate dry conditions prior to 13,000 cal BP.
28,500

Significant cooling Antarctic EDC ice core;


stronger Westerly winds drier at Lago Cardiel.

None

Relatively warmer temperatures in Antarctic


and Pacific slope at 37,000 and 29,000 cal BP;
strong Westerly winds.

8000-6500

13,000-8000

29,000-13,000

ca. 37,000

Age of sediment unit determined by dating organic matter or animal bone (collagen or bioapatite).
Presence of guanaco bones may or may not indicate human presence. If present, human action is indicated by (h) otherwise the bone is used to date sediment.
Pollen information is from La Gruta, La Maria, Los Toldos and La Martita (see Mancini et al.; 2013 for summary and reference s).
3
Time of human presence determined by dating charcoal concentrations or distinct hearths.
4
Information on Antarctic EDC ice core, EPICA Dome C, and Antarctic and Pacific slopes from Kilian and Lamy (2012).

wetter conditions after this date. Pollen data from La Martita Cave 4
also document that this was a period of reduced moisture (Mancini,
1998) as do high levels of Ca, Mg, Na, and K in the upper sediments
of Unit A at LG1 and in Unit 3 at LG3. These dry conditions may
explain the lack of occupation of the Viuda Quenzana and La Gruta
areas until ca. 5583 cal BP (4770  25 14C BP), although more excavations are needed to conrm this absence.
Periods of sediment deposition and non-deposition in the
rockshelters and lagoons at or near La Gruta are compared in
Table 6 with times when there is evidence of human usage of the
rockshelters or when humans were using the nearby area. Of
course, when we deal with archaeological, geological, and palynological information, we are faced with differences in temporal
and spatial resolution. However, we believe important trends can
be discerned from the data we have used. For example, there is a
signicant break in sedimentation at La Gruta 3 between ca. 37,000
and 29,000 cal BP (Unit 1) and ca. 13,000 cal BP (Unit 2), the latter
date agreeing closely with the oldest evidence for human occupation of La Gruta area from 12,799 to 12,049 cal BP based on ages
from Unit A at La Gruta 1 rockshelter. Unit A at La Gruta 1 and Unit 2
at La Gruta 3 appear to have been deposited in the period ca.
13,000e8000 cal BP. Pollen data from La Gruta area, and from other
areas of the Deseado Massif, indicate drier conditions prior to ca.

In contrast, the absence of any evidence for humans in the Viuda


Quenzana and La Gruta areas from about 8800 to 5600 cal BP
(Table 6) suggests a period of drought, a conclusion that is supported by sedimentological evidence from La Gruta Lagoons 1 and 2
and La Barda of dry conditions prior to ca. 6500 cal BP. Low lake
levels at Lago Cardiel, to the west of the area, after ca. 8600 cal BP
(Stine and Stine, 1990), and a change from grass steppe to shrub and
dwarf shrub steppe at La Martita Cave 4 around 8900 cal BP, are
further evidence of dry conditions (Mancini, 1998; Brook et al.,
2013; Mancini et al., 2013).
Evidence of humans at Viuda Quenzana after 5600 cal BP
(guanaco bones showing human action), after 5308 cal BP at La
Martita Cave 4, and at La Gruta after 3600 cal BP (charcoal
scatters), and even more so after 1900 cal BP, corresponds with
evidence of sedimentation in La Gruta lagoons between 6525 and
2150 cal BP indicating wetter conditions. We believe that this
period may have led to some erosion of Unit A at La Gruta 1, and
Unit 2 at La Gruta 3, leading to deposition of Units B and 3 near
the end of the period. Pollen assemblages from Unit 3 at La Gruta
3 and from the basal sediments of Unit B at La Gruta 1 have
relatively high grass pollen percentages suggesting slightly
wetter conditions and possibly more frequent ooding of area
lagoons.

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

18

G.A. Brook et al. / Quaternary International xxx (2014) 1e19

In the last ca. 1500 years, six radiocarbon ages for human
occupation of LG1 (2 ages) and LG3 (4 ages) fall into two distinct
periods: 1372e1271 cal BP and 539e156 cal BP. These periods
predate and postdate the Medieval Climate Anomaly (MCA) that
lasted from ca. 1266 to 539 cal BP, which is considered to have been
a drier interval in southernmost South America (e.g. Stine and Stine,
1990) (Table 6).
7. Conclusions
Sediments in La Gruta area rockshelters and lagoons have provided discontinuous records of climate conditions and periods of
animal and human occupation during the last ca. 37,000 cal BP
(32,650  140 14C BP). Charcoal scatters at La Gruta 1 rockshelter
date the rst human occupation of the area to the late Pleistocene
between 12,799 and 12,049 cal BP (10,845  61e
10,477  56 14C BP) and later to the early Holocene between 9029
and 7760 cal BP (8090  30e7500  250 14C BP). Mylodontidae
bones at La Gruta 3 rockshelter, with dates of 9560  30e
8540  30 14C BP (11,077e9466 cal BP), indicate that the extinct
giant ground sloth continued to use the area even after it had been
occupied by humans. However, as the Mylodontidae dates do not
overlap dated evidence of human occupation during the late
Pleistocene or early Holocene, there is no proof of contact.
The comparison between the pollen assemblages from LG1 and
LG3 and other pollen data for the region, together with information
from the rockshelter sediments, demonstrates that past physicalhuman environments can be reconstructed by analyzing taphonomic processes in areas of human occupation. An important
consideration is the degree of stratigraphic resolution in the rockshelter data but, despite the discontinuities present in these proles palynological richness is similar in both sequences. In addition,
periods with similar pollen assemblages indicate vegetation mosaics in the pollen source area similar to the present mosaic.
The characteristics of sediment units exposed by excavations at
La Gruta 1 and 3 suggest wetter conditions beginning ca. 13,000
cal BP and lasting until ca. 8000 cal BP (10,845  61e
7500  250 14C BP) and then again from ca. 6500 to 1300 cal BP
(5690  35e1452  38 14C BP for La Barda and LG1, respectively).
The rst of these intervals includes the late Pleistocene and early
Holocene periods of human occupation suggesting humans utilized
the area during times of increased moisture. Pollen data are available for the rst of these periods and conrm the sediment data
indicating more moisture at these times. Lacustrine silts and clays
in La Barda, and La Gruta Lagoons 1 and 3 also indicate wetter
conditions in the period 6500e1300 cal BP and in addition provide
evidence of an arid interval prior to about 6500 cal BP
(5690  35 14C BP). This may explain why there is no evidence of
humans between ca. 7760 and 3830 cal BP (7500  250 and
3487  38 14C BP) who could have moved their activities to nearby
areas with a more reliable water supply, such as the Chico River
basin. However, there is clear evidence of human occupation of the
Viuda Quenzana area after 5580 cal BP and at La Gruta around 3800
and even more so after 1900 cal B.P. The ndings in La Gruta area
show that Mylodontidae was probably present in the southern
Deseado Massif after the rst humans arrived as is also suggested
by data from the Piedra Museo archaeological locality to the north.
Based on evidence from southern Patagonia, Mylodontidae became
extinct soon afterwards.
Acknowledgements
Funding was provided by PIP (CONICET) 0356, Cooperation
project CONICET-NSF (Res.1838/13; 20132015), and the Franklin
College of the University of Georgia. The University of Arizona and

University of Georgia radiocarbon dating laboratories assisted with


radiocarbon dating. We thank paleontologists Eduardo Tonni,
Scillato Yane, Federico Agnolin and also Sergio Bogan for their
considerable help in identifying the Mylodontidae remains. We
acknowledge the assistance of the managements of the Triton
Mining Company S.A. and Piedra Grande mines, and appreciate the
support and help of the Gobernador Gregores authorities (sr. Pablo
Ramrez and Marcelo Cebeira). We thank, in particular, Carlos Baetti
for his considerable assistance and for providing photographs of the
historic 1983 ood at La Gruta Lagoon 1, taken by the Duncan
family, from San Julin, the previous owners of Estancia La Gruta.
We also wish to thank Lucas Vetrisano for digging the Entrance Pit
at La Gruta 3. We also wish to thank geologist Claudio Iglesias, from
Piedra Grande S.A. who helped us with the geology and specic
rock types of La Gruta area. To all the people who helped with eld
work, we owe our deepest gratitude.
References
Aguerre, A.M., 1982. Informe preliminar de la excavacin de la Cueva 4 de La
Martita, departamento de Magallanes, Santa Cruz. Paper presented to the VII
Congreso Nacional de Arqueologa, San Luis, Argentina.
Aguerre, A.M., 2003. La Martita: Ocupaciones de 8000 aos en la Cueva 4. In:
Aguerre, A. (Ed.), Arqueologa y Paleoambiente en la Patagonia Santacrucea
Argentina. Ediciones del autor, Buenos Aires, pp. 29e61.
Bamonte, F.P., Mancini, M.V., 2011. Palaeoenvironmental changes since Pleistocenee
Holocene transition: pollen analysis from a wetland in southwestern Patagonia
(Argentina). Review of Palaeobotany and Palynology 165, 103e110.
Bargo, M.S., Vizcano, S.F., 2008. Paleobiology of Pleistocene ground sloths (Xenarthra, Tardigrada): biomechanics, morphogeometry and ecomorphology applied
to the masticatory apparatus. Ameghiniana 45, 175e196.
Bargo, M.S., De Iuliis, G., Vizcano, S.F., 2006a. Hypsodonty in Pleistocene ground
sloths. Acta Paleontologica Polonica 51, 53e61.
Bargo, M.S., Toledo, N., Vizcano, S.F., 2006b. Muzzle of south American ground
sloths (Xenarthra, Tardigrada). Journal of Morphology 267, 248e263.
Barnosky, A.D., Lindsey, E.L., 2010. Timing of Quaternary megafaunal extinction in
South America in relation to human arrival and climate change. Quaternary
International 217, 10e29.
Belardi, J.B., Tiberi, P., Stern, C., Sunico, A., 2006. Al Este del Cerro Pampa: ampliacin
del rea de disponibilidad de obsidiana de la Pampa del Asador (Provincia de
Santa Cruz). Intersecciones en Antropologa 2006 (7), 27e36 (online).
Bird, J., 1988. Travels and Archaeology in South Chile. University of Iowa Press, Iowa
City.
Blott, S., Pye, K., 2001. GRADISTAT: a grain size distribution and statistics package
for the analysis of unconsolidated sediments. Earth Surface Processes and
Landforms 26, 1237e1248.
Borrero, L.A., 2001. El poblamiento de la Patagonia. Toldos, milodones y volcanes.
Emec, Buenos Aires.
Borrero, L.A., 2012. The human colonization of the high Andes and southern South
America during cold pulses of the late Pleistocene. In: Eren, M.I. (Ed.), Huntergatherer behaviour: Human response during the Younger Dryas. Left Coast
Press, Inc, Walnut Creek, California, pp. 57e78.
Borrero, L., Lanata, J., Borella, F., 1997. La extincin de la megafauna en la Patagonia.
Anales del Instituto de la Patagonia. Serie Ciencias Humanas 25, 89e102.
Borrero, L.A., Martin, F., 2008. A reinterpretation of the Pleistocene human and
faunal association at Las Buitreras Cave, Santa Cruz, Argentina. Quaternary
Science Reviews 27, 2509e2515.
Brandoni, D., Ferrero, B.S., Brunetto, E., 2010. Mylodontidae darwini Owen (Xenarthra, Mylodontidaetinae) from the Late Pleistocene of Mesopotamia, Argentina,
with remarks of individual variability, paleobiology, paleobiogeography, and
paleoenvironment. Journal of Vertebrate Paleontology 30 (5), 1547e1558.
Brook, G.A., Mancini, M.V., Franco, N.V., Bamonte, F., Ambrstolo, P., 2013. An examination of possible relationships between paleoenvironmental conditions
during the Pleistocene-Holocene transition and human occupation of southern
Patagonia (Argentina) east of the Andes, between 46 and 52 S. Quaternary
International 305, 104e118.
Coronato, A., Salemme, M., Rabasso, J., 1999. Palaeoenvironmental conditions during the early peopling of southernmost South America (Late Glacial e Early
Holocene, 14e8 ka B.P.). Quaternary International 53/54, 77e92.
De Porras, M.E., Maldonado, A., Abarza, A.M., Crdenas, M.L., Francois, J.-P., Martel
Cea, A., Stern, C.R., Mendez, C., Reyes, O., 2012. Postglacial vegetation, re and
climatic dynamics at central Chilean Patagonia (Lake Shaman, 44 S). Quaternary Science Reviews 50, 71e85.
Durn, V., Gil, A., Neme, G., Gasco, A., 2003. El Verano: ocupaciones de 8900 aos en
la Cueva 1 (Santa Cruz, Argentina). In: Aguerre, A. (Ed.), Arqueologa y Paleoambiente en la Patagonia Santacrucea Argentina. Ediciones del autor, Buenos
Aires, pp. 93e120.
Farrand, W.R., 1975a. Sediment analysis of a prehistoric rockshelter. Quaternary
Research 5, 1e21.

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

G.A. Brook et al. / Quaternary International xxx (2014) 1e19


Farrand, W.R., 1975b. Analysis of the Abri Pataud sediments. In: Movius Jr., H.L. (Ed.),
Excavation of the Abri Pataud, Les Eyzies, Dordogne. Peabody Museum of
Archaeology and Ethnology, Harvard University, Cambridge, MA, pp. 27e68.
Farrand, W.R., 2001. Sediments and stratigraphy in rockshelters and caves: a personal perspective on principles and pragmatics. Geoarchaeology: An International Journal 16 (5), 537e557.
Folk, R.L., Ward, W.C., 1957. Brazos River bar: a study in the signicance of grain size
parameters. Journal of Sedimentary Petrology 27, 3e26.
Franco, N.V., Otaola, C., Cardillo, M., 2007. Resultados de los trabajos exploratorios
realizados en la margen norte del ro Santa Cruz (provincia de Santa Cruz,
Argentina). In: Morillo, F., Martinic, M., Prieto, A., Bahamonde, G. (Eds.),
Arqueologa de Fuego-Patagonia. Levantando piedras, desenterrando huesos.
y develando arcanos. Ediciones CEQUA. Punta Arenas, Chile, pp. 541e553.
Franco, N.V., Martucci, M., Ambrstolo, P., Brook, G.A., Mancini, M.V., Cirigliano, N.,
2010. Ocupaciones humanas correspondientes a la transicin PleistocenoHoloceno al sur del Macizo del Deseado: el rea de La Gruta (provincia de
Santa Cruz, Argentina). Relaciones de la Sociedad Argentina de Antropologa
XXXV, 301e308.
Franco, N.V., Ambrstolo, P., Martucci, M., Brook, G.A., Mancini, M.V., Cirigliano, N.,
2011. Early human occupation in the southern part of the deseado massif
(Patagonia, Argentina). Current Research in the Pleistocene 27, 13e16.
Franco, N.V., Ambrstolo, P., Cirigliano, N., 2012. Disponibilidad de materias primas
lticas silceas en el extremo sur del Macizo del Deseado: los casos de La Gruta y
Viuda Quenzana. Magallania 40 (1), 279e286.
Franco, N.V., Ambrstolo, P., Acevedo, A., Cirigliano, N., Vommaro, M., 2013. Prospecciones en el sur del Macizo del Deseado (provincia de Santa Cruz). Los casos
de La Gruta y Viuda Quenzana. In: Zangrando, A.F., Barberena, R., Gil, A.,
Neme, G., Giardina, M., Luna, L., Otaola, C., Paulides, S., Salgn, L., Tvoli, A. (Eds.),
Tendencias terico-metodolgicas y casos de estudio en la Arqueologa de la
Patagonia. Museo de Historia Natural de San Rafael. Altuna Impresores, Buenos
Aires, pp. 371e378.
Heusser, C.J., 1995. Three Late Quaternary pollen diagrams from Southern Patagonia
and their palaeoecological implications. Palaeogeography, Palaeoclimatology,
Paleoecology 118, 1e24.
Heusser, C.J., Rabassa, J., 1987. Cold climate episode of Younger Dryas age in Tierra
del Fuego. Nature 328, 609e611.
Heusser, C.J., Streeter, S.S., 1980. A temperature and precipitation record of the past
16000 years in southern Chile. Science 210 (4476), 1345e1347.
Heusser, C.J., Borrero, L.A., Lanata, J.L., 1992. Late Glacial vegetation at Cueva del
Mylodon. Anales del Instituto de la Patagonia (Serie Ciencias Naturales) 21, 97e
102.
Hogg, A.G., Hua, Q., Blackwell, P.G., Buck, C.E., Guildersno, T.P., Heaton, T.J., Niu, M.,
Palmer, J.G., Reimer, P.J., Reimer, R.W., Turney, C.S.M., Zimmerman, S., 2013.
Radiocarbon 55 (4), 1889e1903.
Indorante, S.J., Hammer, R.D., Koenig, P.G., Follmer, L.R., 1990. Particle-size analysis
by a modied pipette procedure. Soil Science Society of America Journal 54,
560e563.
Jackson, D., 2002. Los instrumentos ltiocos de los primeros cazadores de Tierra del
Fuego. In: Ediciones de la Direccin de Bibliotecas. Archivos y Museos, Santiago
de Chile, Chile.
Kilian, R., Lamy, F., 2012. A review of Glacial and Holocene paleoclimate records
from southernmost Patagonia (49e55 S). Quaternary Science Reviews 53, 1e
23.
Long, A., Martin, P.S., Lagiglia, H.A., 1998. Ground sloth extinction and human
occupation at Gruta del Indio, Argentina. Radiocarbon 40 (2), 693e700.
Mancini, M.V., 1998. Vegetational changes during Holocene in the extra-Andean
Patagonia, Santa Cruz province, Argentina. Palaeogeography, Palaeoclimatology and Palaeoecology 138 (1e4), 207e219.
Mancini, M.V., 2009. Holocene vegetation and climate changes from a peat pollen
record of the forest-steppe ecotone, Southwest of Patagonia (Argentina). Quaternary Science Reviews 28, 1490e1497.
Mancini, M.V., de Porras, M.E., Bamonte, F.P., 2012. Southernmost south america
steppes: vegetation and its modern pollen assemblages representation. In:
Germanno, Denise M. (Ed.), Steppe Ecosystems: Dynamics, Land Use and Conservation, pp. 141e156. Edit. Nova, 218 pp.
Mancini, M.V., Franco, N.V., Brook, G.A., 2013. Palaeoenvironment and early human
occupation of southernmost South America (South Patagonia, Argentina).
Quaternary International 299, 13e22.
Markgraf, V., 1985. Late Pleistocene faunal extinction in southern Patagonia. Science
228, 1110e1112.
Martin, F.M., San Romn, M., Morello, F., Todisco, D., Prevosti, F.J., Borrero, L., 2013.
Land of ground sloths: recent research at Cueva Chica, ltima Esperanza, Chile.
Quaternary International 305, 56e66.
Massone, M., 1987. Los cazadores paleoindios de Tres Arroyos (Tierra del Fuego).
Anales del Instituto de la Patagonia (Serie Ciencias Sociales) 17, 47e60.

19

Massone, M., 1996. Hombre temprano y paleoambiente en la regin de Magallanes:


evaluacin crtica y perspectivas. Anales del Instituto de la Patagonia (Serie
Ciencias Sociales) 24, 81e98.
Massone, M., Prieto, A., 2004. Evaluacin de la Modalidad Cultural Fell 1 en Magallanes. Chungara 36, 305e315.
Mehlich, A., 1953. Determination of P, Ca, Mg, K, Na, and NH4. North Carolina Soil
Test Division (Mimeo 1953).
Mengoni, G.L., 1987. Modicaciones culturales y animales en los huesos de los
niveles inferiores del sitio Tres Arroyos 1 (Tierra del Fuego). Anales del Instituto
de la Patagonia (Serie Ciencias Sociales) 17, 61e66.
Miotti, L., Salemme, M., 2003. When Patagonia was colonized: people mobility at
high latitudes during Pleistocene/Holocene transition. Quaternary International
109-110, 95e111.
Miotti, L., Salemme, M., 2004. Peopling, mobility and territorios between the
hunter-gatherers populations in Patagonia. Complutum 15, 177e206.
Miotti, L., Vzquez, M., Hermo, D., 1999. Piedra Museo: Un yamnagoo pleistocnico
de los colonizadores de la Meseta de Santa Cruz: el estudio de la Arqueofauna.
Soplando en el viento. Actas de las Terceras Jornadas de Arqueologa de la
Patagonia. S. C. de Bariloche.
Moore, D.M., 1978. Post-glacial vegetation in the South Patagonian territory of the
giant ground sloth, Mylodon. Botanical Journal of the Linnean Society 77 (3),
177e178.
Nami, H., 1985-86. Excavacin arqueolgica y hallazgo de una punta de proyectil
Fell I en la cueva del Medio. Seno de ltima Esperanza, Chile. Anales del
Instituto de la Patagonia (Serie Ciencias Sociales) 16, 103e109.
Nami, H., Nakamura, T., 1995. Cronologa radiocarbnica con AMS sobre muestras
de huesos procedentes del sitio Cueva del Medio (ltima Esperanza, Chile).
Anales del Instituto de la Patagonia (Serie Ciencias Sociales) 23, 125e133.
Nordenskjld, E., 1996 [1900]. Observaciones y descubrimientos en cuevas de Ultima Esperanza en Patagonia Occidental. Anales del Instituto de la Patagonia,
Serie Ciencias Humanas 24, 99e124.
Paez, M.M., Prieto, A.R., Mancini, M.V., 1999. Fossil pollen from Los Toldos locality: a
record of the Late-glacial transition in the Extra-Andean Patagonia. Quaternary
International 53e54, 69e75.
Panza, J.L., Marin, G., 1998. Hoja Geolgica 4969-I Gobernador GregoresProvincia
de Santa Cruz. In: Boletn 239. SEGEMAR, Buenos Aires. Argentina.
Paunero, R., 2009. In: El arte rupestre milenario de Estancia La Mara, Meseta
Central de Santa Cruz. Estudio Denis, La Plata.
Paunero, R.S., Frank, A.D., Skarbun, F., Rosales, G., Cueto, M., Zapata, G., Paunero, M.,
Lunazzi, N., Del Giorgio, M., 2007. Investigaciones arqueolgicas en sitio Casa
del Minero 1, Estancia La Mara, Meseta Central de Santa Cruz. In: Morello, F.,
Martinic, M., Prieto, A., Bahamonde, G. (Eds.), Arqueologa de Fuego-Patagonia.
Levantando piedras, desenterrando huesos y develando arcanos. Ediciones
CEQUA. Punta Arenas, Chile, pp. 577e588.
Prevosti, F.J., Martin, F.M., 2013. Paleoecology of the mammalian predator guild of
Southern Patagonia during the latest Pleistocene: ecomorphology, stable isotopes, and taphonomy. Quaternary International 305, 74e84.
Smith, M., Veth, P., Hiscock, P., Wallis, L.A., 2005. Global deserts in perspective. En
Desert peoples. Archaeological perspectives. In: Veth, P., Smith, M., Hiscock, P.
(Eds.), Desert People. Archaeological Perspectives. Blackwell Publishing Ltd,
Pondicherry, India, pp. 1e13.
Sottile, G.D., Bamonte, F.P., Mancini, M.V., Bianchi, M.M., 2012. Insights into Holocene vegetation and climate changes at the Southeastern side of the Andes:
Nothofagus Forest and Patagonian Steppe re records. The Holocene 22 (11),
1201e1214.
Stern, C.R., 2000. Sources of obsidian artifacts from the Pali Aike, Fells Cave and
Caadn La Leona archaeological sites in southernmost Patagonia. In: Desde el
Pas de los Gigantes. Perspectivas arqueolgicas en Patagonia. Universidad
Nacional de la Patagonia Austral, Ro Gallegos, pp. 43e55.
Stine, S., Stine, M., 1990. A record from Lake Cardiel of climate in southern South
America. Nature 345, 705e708.
Stuiver, M., Reimer, P.J., 1993. Extended 14C database and revised CALIB radiocarbon
calibration program. Radiocarbon 35, 215e230.
Tonello, M.S., Mancini, M.V., Sepp, H., 2009. Quantitative reconstruction of Holocene
precipitation changes in Southern Patagonia. Quaternary Research 72 (3), 410e420.
U.S. EPA, 2014. Method 6010C. In: SW-846 On-line Test Methods for Evaluating
Solid Waste, Physical/Chemical Methods. http://www.epa.gov/osw/hazard/
testmethods/sw846/online/index.htm.
Veth, P., 2005. Cycles of aridity and human mobility: risk minimization among late
pleistocene Foragers of the Western Desert, Australia. In: Veth, P., Smith, M.,
Hiscock, P. (Eds.), Desert People Archaeological Perspectives. Blackwell Publishing Ltd, Pondicherry, India, pp. 100e115.
Villa-Martnez, R., Moreno, P.I., 2007. Pollen evidence for variations in the southern
margin of the westerly winds in SW Patagonia over the last 12,600 years.
Quaternary Research 68, 400e409.

Please cite this article in press as: Brook, G.A., et al., Evidence of the earliest humans in the Southern Deseado Massif (Patagonia, Argentina),
Mylodontidae, and changes in water availability, Quaternary International (2014), http://dx.doi.org/10.1016/j.quaint.2014.04.022

Quaternary International xxx (2014) 1e8

Contents lists available at ScienceDirect

Quaternary International
journal homepage: www.elsevier.com/locate/quaint

Moving: Hunter-gatherers and the cultural geography of South


America
Luis Alberto Borrero
CONICET-IMHICIHU, Saavedra 15, Piso 5, 1083ACA Buenos Aires, Argentina

a r t i c l e i n f o

a b s t r a c t

Article history:
Available online xxx

The conditions under which the process of human colonization of South America took place are discussed. The modes of acquisition of environmental knowledge, as a way to construct a cultural geography, are also considered. An example concerning the peopling of the forests, particularly in Northwest
South America, and the role of plants in the early stages of colonization is also offered. Finally the signicance of non-utilitarian items, exchange, and empty lands for our understanding of the process of
peopling is discussed.
2014 Elsevier Ltd and INQUA. All rights reserved.

Keywords:
Colonization
South America
Plants
Knowledge

1. Introduction
Even when there is no consensus about when and how the
process of peopling of South America started, the available evidence indicates that ecologically disparate regions of the continent
were already occupied around 10,000 BP (Politis, 1999; Dillehay,
2000; Aceituno et al., 2013). This was a process that surely
involved generalist hunter-gatherers with the necessary exibility
to exploit different niches. There is archaeological evidence of
diverse lithic industries, use of large and small terrestrial and
marine vertebrates, and intense exploitation of plant resources
(Stahl, 1996; Dillehay, 2000; Ranere and Lpez, 2007). At the same
time, the existence of this variety of adaptations requires a long
previous history of peopling. No matter how fast was the process
of human peopling, several generations of people interacting with
the environments, and with the local climates, would be needed to
be successful in so many regions. These people have to understand
the new environment and then transform it as a result of its
exploitation.
The variety of habitats exploited ca. 10,000 BP also suggests that
the history of the human expansion into South America was not
simple, and that a number of theoretical and practical issues should
be considered. The situation is of course similar to that of the colonization of other regions of the world. From a theoretical point of
view what is implied is that the net diffusion through time was
simple a by-product of how people lived in landscapes (Denham
et al., 2009: 29), in other words an exaptation (see Gamble, 1994).

E-mail address: laborrero2003@yahoo.com.

If this explanation is valid, then there is no requirement of major


migrations, be it fast or slow, to explain the displacement of people.
On the other hand, practical issues fall within the purview of what
can be called a taphonomic approach to the archaeology of peopling.
In the rst place it includes what I call Regional taphonomy, that is
a concern for the distribution of preservational pockets in the
landscape and the study of the mechanisms that accumulate and
preserve materials (Borrero, 2001). The construction of a continental scale taphonomy is a difcult task, one that can only be
delineated at this time. The basic idea is to apply this approach at the
same geographical scale at which archaeological projects work. The
goal is a better denition of the archaeological problems implicated
in the processes of exploration and colonization. A rst distinction is
between large environmental patches, as can be dened for the Late
Pleistocene (Clapperton, 1993), and a relatively sharp denition of
the relevant habitats for the rst inhabitants within those patches.
These can be dened on the basis of paleoecological research,
particularly the paleodistribution of corridors and other biogeographic features. Variation along a number of taphonomically
relevant properties can be examined. Among other measures, the
proportion of space covered by different classes of soils constitute a
rst approximation to understand differences in bone preservation
among patches, while charts of the impact of erosion mark differences in the feasibility of burial and general visibility of the
archaeological record. A ranking of past habitats in terms of
archaeologically relevant properties should be the main result. For
example, the evidence showing that large parts of the Pacic coast of
South America were affected by the action of tsunamis is relevant for
our assessment of the early exploration of the coastal habitats
(Lpez-Castao and Cano-Echeverra, 2012: 49).

http://dx.doi.org/10.1016/j.quaint.2014.03.011
1040-6182/ 2014 Elsevier Ltd and INQUA. All rights reserved.

Please cite this article in press as: Borrero, L.A., Moving: Hunter-gatherers and the cultural geography of South America, Quaternary
International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.011

L.A. Borrero / Quaternary International xxx (2014) 1e8

The amount of knowledge of the environment available to the


early colonizers can be inferred from the archaeological record. For
example, after examining the evidence from early coastal sites in
Peru, Dan Sandweiss was able to conclude that people knew how
to exploit the sea when they rst arrived in Western South America,
or shortly thereafter (Sandweiss, 2008: 153). Information to
discuss this at the continental scale is not available, but some cases
can be explored. The rst product of this approach is a re-reading of
the archaeological record in terms of evidences of knowledge of the
environment by the rst inhabitants. Only a very crude approximation to these issues can be considered here, where we will
specically discuss issues related with the archaeology of tropical
forests, of empty lands and the importance of non-utilitarian items.
2. Exploration
An ecological model of this process may be useful in selecting
the relevant data. I have previously used such a model to organize
the archaeological information from Fuego-Patagonia (Borrero,
1989e1990, 1989). This model contemplates the human exploration of new lands, sometimes followed by colonization and effective occupation. The reason to call one stage of this process
colonization is that it is difcult to view a group of explorers as
cut off from their original population (Rogers, 1990). The biological
viability of those explorers that will allow them to be colonizers is
based on the fact that ties with their mother group are not shut off.
One of the main properties of this model is that it does not require
constant southward movement, but only a slow multidirectional
ow of people. In some way, this is convergent with results of
human morphological studies that indicate that the peopling process was probably the result of multiple discrete expansions of
highly variable founder populations (Delgado-Burbano, 2012: 35).
Discussing the early archaeology of North America, Hofman wrote
that many times the repeated use of specic high-quality lithic
sources led to believe that their long-term pattern of land use
should have resulted in lithic distributional patterns suggesting
one-way movement, even if people moved in complex patterns
(Hofman, 2003: 234).
The mechanisms behind movement probably included the
gradual extension of hunting ranges, the ssion of bands, the search
for high quality raw material sources, and perhaps also starvation,
curiosity, and other causes, principal among them the simple act of
living within a variable home range (Anderson and Gillam, 2000;
Belovsky, 1987; McGhee, 1997: 125e126). Problems in the home
territory may also be a cause for movement, as recorded in the
classic ethnographic example of the 19th Century Inuit migration
(Mary-Rousselire, 2008 [1980]). In general terms, Kelly described
the situation of expansion as one of giving up a known environment for an unknown environment (Kelly, 1999: 124). It is true
that hunter-gatherers surely bring with them a variety of strategies
and technologies useful for a number of circumstances, but this
does not implies that people never enter unknown territory
(Randall and Hollenbach, 2007: 220).
The availability of hierarchically ordered space, and the structure of critical resources should have directed people in different
directions, not necessarily lling all ground behind. Places with
fauna that lack anti-predatory behaviors were probably initially
favored, even when most published studies suggest that these behaviors were probably rapidly learned (Berger et al., 2001). A strong
negative impact on the success of explorers could be the result of
the prey increasing vigilance or improving its escape abilities.
For this and other reasons, the resulting distribution of people
should be discontinuous, leaving many empty zones and with some
differences between settling-in and on-the-move places. The
visibility of those places should be very different, and it can be

maintained that most of the discovered early archaeological sites


correspond to the rst class. The usual trend toward the study of
large sites goes against chances of nding sites related with an
exploration stage.
The criteria to nd and recognize the rst stages in the process
of exploration and colonization of any region are not completely
understood. Generally speaking, archaeological markers that signal
lack of local knowledge are useful, because they are indicative of
partial familiarity with the local geography. In another level, they
also mark the possibility of maladaptations, suggesting that local
extinctions (extirpations) and cultural failures may happen. A
recent review of the limited evidence for the earlier human remains
in America showed that earlier people were living a life with a
signicant amount of risk, and that stress on Paleoamerican females makes it unlikely that the population of the rst Americans
could have grown rapidly (Chatters, 2010: 67). The result at a
supra-regional scale should be spatial discontinuity of the human
settlement (Butzer, 1988). Similar situations are modeled by the
point and arrow pattern proposed by Rockman, in which there is
movement in which colonizers stream from known areas to new
areas, leaving the areas in between uncolonized (Rockman, 2003:
9). I have reiteratively sustained that early settlers need not have a
perfect adjustment to their environments. For example, the cases of
the Holocene sites Tnel and Imiwaia in Tierra del Fuego (Fig. 1) are
good examples of places where the knowledge of the local resources appear not to be high for the rst inhabitants (Piana et al.,
2012), a situation that contrasts with later occupations that indicate
a detailed knowledge of the local resources (Orquera and Piana,
2009).
The potential markers of the degree of familiarity with the local
resources are varied, including evidences of sub-optimal use of the
available resources (Muscio, 2001). Exploration refers to the initial
radiation of humans to new empty land (see Borrero, 1994-1995).
Less resistance routes are usually implicated and most of the
settling-in places are probably widely separated. Undoubtedlly, the
visibility of relevant materials should be low, since sub-optimal
places probably were not reoccupied. The basic criteria to recognize these sites include chronological precedence, in other words
the older sites or older archaeological strata within a region are
candidates. Application of this criterion is in no way restricted to
the Late Pleistocene, but to the older evidence in any given habitat
or region. The presence of few remains should testify to exploration
stage occupations, many times at sub-optimal locations. Identication of the substrate on which the older occupation rests is also
informative. For example in large sections of northeast Tierra del
Fuego, the older substratum is slightly older than 4000 radiocarbon
years. Any occupation around that age which is resting on that
substratum is a candidate for an exploration stage representative.
Similar situations with dates immediately after deglaciation exist
along the Andean Cordillera.
More specically, limited redundancy in the early occupations
and the existence of occupational gaps indicating discontinuity in
human installation, with cases of alternate use by carnivores and
humans, are also expected. Trans-generational time frames should
be usually implicated. Other expectations include use of abundant
local raw materials, independently of its quality. Moreover, Franco
studied the criteria to recognize an exploration stage using lithic
artifacts. She expects tools not to be broken, as they should be
expediently made on local rocks. Long-cutting edges should be
dominant and the few cases of exotic rocks are to be understood in
the context of personal gear (Civalero and Franco, 2003; Franco,
2003). All these expectations were met in her analysis of the
early Patagonian assemblages. Importantly, she concluded that
versatility (sensu Nelson, 1991) is adequate for the task, particularly
bifacial tools with high transportability (Kelly, 1988). A number of

Please cite this article in press as: Borrero, L.A., Moving: Hunter-gatherers and the cultural geography of South America, Quaternary
International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.011

L.A. Borrero / Quaternary International xxx (2014) 1e8

Fig. 1. South America and location of archaeological sites mentioned in the text. A Porce Basin, B Amazon Basin, C Rio de la Plata Basin, D Simpson Basin.

studies in different cultural settings of the world recently suggested


that levallois technology was adequate for wide ranging movements in relatively unknown territory (White and Pettitt, 2011: 75).
This is relevant because levallois technology was recognized in
South America, and its distribution and signicance is just beginning to be understood (Nami, 1992; Franco, 2004; Morello, 2005).
Generally speaking, the early stages of exploration promote the
conditions under which there is a role for exaptations (Borrero and
Borrazzo, 2013). The main reason is that in environments under
colonization there are new needs and also unknown or poorly
known resources. Hunter-gatherers have ways to deal with unexpected situations: for example Binford describes the skills of the
Umialuk among the Nunamiut, who knows how to use knowledge
and has great knowledge depth regarding the long-term behavior
of animals e what might be called the regularity in their erratic
behavior (Binford, 1991: 55). This knowledge simply covers the
behavioral range of a known species. It is a different story to deal
with new species for which knowledge is incomplete. Importantly,
there is always the risk of applying preconceived and perhaps
faulty models to a landscape (Meltzer, 2009: 372). This is one of
the reasons why colonizers of diverse lands probably were not
specialists. This marks an important difculty with Kelly and Todds
model of colonization, which states that the rst inhabitants of a
region has no need to occupy new niches (1988: 235). This can be
sustained for North America, but only by restricting their

movements to the Andean region it was possible for the explorers


of South America to remain in the same niche. The available evidence on the distribution of early sites suggests that a variety of
niches were exploited by the end of the Pleistocene (Roosevelt
et al., 1996; Borrero, 2004; Aceituno et al., 2013.). Another reason
for the early explorers not to be specialists is that on all accounts
the dependency on meat, particularly lean meat, known as Rabbit
starvation, should be avoided by including carbohydrates in the
diet (Speth and Spielmann, 1983). Early explorers should have wide
generalist diets, which perhaps should differ markedly from diets
of posterior inhabitants. Although the bioanthropological record for
early population is meager, these changes should be detectable in
stable isotopes values, and perhaps in other bioanthropological
markers (Guichn, 1994; Chatters, 2010).
3. Peopling South America
Early explorers of South America were probably using varied
criteria in order to rank potentially attractive habitats, in relation
with their previous knowledge and technium (Arthur, 2009;
Borrero et al., 2013), which are the basic tools used to transform
the geography. At the landscape scale, habitats are larger units than
those that Optimal Foraging Theory usually treat as a patch,
sometimes similar to what Beaton (1991) called megapatches.
Quality and availability of raw materials and other basic subsistence

Please cite this article in press as: Borrero, L.A., Moving: Hunter-gatherers and the cultural geography of South America, Quaternary
International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.011

L.A. Borrero / Quaternary International xxx (2014) 1e8

needs were probably paramount in deciding which places were


more interesting to be settled, but probably were not the only
criteria. Minimally, distance to their original populations and the
location of neighbors must have been important. For example, the
Pacic coastal habitats are largely recognized as highly productive
and homogeneous along long distances (Dixon, 2001; Erlandson,
2001; Kelly, 2003a), and can be considered as a series of megapatches (sensu Beaton, 1991), where people simply kept with an
adaptation that they knew (Kelly, 2003a: 139). Miotti considered
the alternative of a process of peopling following the Atlantic coast
of South America which could have also involved homogeneous
megapatches (Miotti, 2003: 163e164). However, there is no evidence of strong selectivity for those coastal environments during
early times. The borders of patches or megapatches are attractive
spots, usually associated with access to a larger set of subsistence
resources. On that basis Martino et al. conclude that, when a
resource diminishes in one habitat, they can have quick access to
equivalent or alternative resources from another habitat; and this
characteristic may favor dispersion (Martino et al., 2007: 6). This
suggestion ignores the fact that the process of colonization could
not be accomplished simply by appropriation of what is available,
but that it is mediated by a process of reconnaissance of the existent
resources and the ways of exploiting them. The archaeological record is also a record of the increasing knowledge of the environment and its resources through time, which in turn is associated
with its transformation into a cultural geography. Corridors in NW
South America produced by deglaciation and volcanic activity
probably facilitated the access to the forests (Lpez-Castao and
Cano-Echeverra, 2012), a process that necessary implicates successive changes in the list of exploitable resources. The construction of habitat implicates both reversible and irreversible changes
which will have differential archaeological results and signals.
Discussions about the efciency of the adaptations exist, but they
are not usually very useful. For example, Riches (1982) suggested
that individual hunting was more efcient than herd hunting, but
there is a difculty in comparing the efciency of different strategies (Aschero and Martinez, 2001: 236e237), and both classes of
strategies may prove adequate or even optimal under different sets
of conditions. The point is that they will produce a quite distinct
archaeological signal on the landscape and a different effect on the
prey population. In the rst place, efcient strategies are more
visible, simply because they are associated with higher redundancy
in the occupation.
3.1. Forests
More specic examples are provided by the available information for the exploration of the tropical forests of Colombia or Brazil.
The timing of the early exploration of forests is an issue that goes
beyond the case of South America. The Amazonian evidence runs
counter to the principle asserting that only the possession of metal
tools allows humans to colonize forested environments (see Politis
and Gamble, 1994). However, one thing is to recognize that humans
are fully capable of coping with forested environments, and
another to maintain that the process of expansion to the forests was
without difculties. It was shown that Despite the high diversity of
species useful plants are few (Aceituno-Bocanegra and Castillo
Espita, 2005: 4). Quoting archaeological observations by Gnecco
and ethnoarchaeological evidence provided by Politis, Scheinsohn
sustains that some researchers support the idea that. people
created their own patches of resources in order to increase their
effectiveness in the environment (Scheinsohn, 2003: 345). This is
true in general, of course, but it is probably not something that early
explorers achieve immediately. On the basis of evidence collected
at the Porce valley, Colombia, it was sustained the existence of early

Holocene logistical camps from which the hunter gatherers set out
for other zones of the basin in order to obtain resources and information (Aceituno-Bocanegra and Castillo Espita, 2005: 5). On
that basis it is claimed that the lower levels of those sites resulted
from explorers occupying an unknown land. Certainly, the presence
of stone axes, cutting and scraping tools and quern stones, and the
low levels of disturbance of the forest support this claim. Some time
for adaptation was required before hunter-gatherers possess the
necessary knowledge and skills to modify their environment according to their needs, which in the case of the Porce valley
included increasing diversity of lithic tools and the construction of
stone oorings (Aceituno-Bocanegra and Castillo Espita, 2005: 6).
One problem with forests is that plant-related information and
non-organic resources, such as lithics, may have very low transferable value from habitat to habitat (Rockman, 2003: 19). This
should have been especially pressing in environments such as the
Amazon basin, where lithics are very scarce. Also, the lack of
topographic relief makes navigation more difcult, retarding the
process of learning the intricacies of the landscape (Kelly, 2003b:
49, 54). Moreover, hunter-gatherers cannot endanger themselves
by collecting and consuming unknown wild plants. An extreme
example is provided by the process of acquiring the knowledge to
recognize and eventually process toxic plants, including for
example Lonchacarpus nicou which was used to hunt sh in
ethnographic times (Crdenas and Politis, 2000: 59). It is expected
that this plants are not incorporated early in the process of gathering information on the environment, and for some the use of
toxic plants for subsistence signals the existence of some kind of
stress (OConnell and Allen, 2012).
In tropical forests, we have the important evidence obtained by
Gnecco and Mora for the early Holocene (Gnecco and Mora, 1997;
Gnecco, 2000). Sites San Isidro and Pea Roja, Colombia were
occupied ca. 10,000-9000 BP, and the archaeological remains suggest a non-especialized extractive technology (Mora and Gnecco,
2003: 275), although the stone axes with side notches and the
mortar-like stones of Pea Roja may indicate tree-falling and nutcracking respectively (Oliver, 2008: 202), and nely made stone
hoes are recorded at several early Holocene sites in NW South
America (Piperno, 2011a:S460). The inhabitants of Pea Roja
arrived before 9000 BP within a rain forest context. Some changes
through time were recorded, including reductions in the abundance of charcoal concomitant with the introduction of squashes
(Cucurbita spp.) (Mora and Gnecco, 2003: 276). Several species of
palms with possible economic value are also recorded. The evidence at San Isidro, located at w1600 m asl, clearly indicates that
some previous knowledge of the area existed. More than 65000
lithic artifacts were found, suggesting that this place was redundantly used. The variety of raw materials, some of them from
distant sources, indicates that formal exploration took place before
the intense utilization of that place. In that sense, exploitation of
the very small, buried obsidian ows in the valley of Popayn/
underscores/. a detailed territorial knowledge (Gnecco, 2003a:
18). There is not a well preserved faunal record, but there are
fascinating evidences of the consumption of a variety of plants e
charred seeds of Persea spp. and Erythrina and starch grains from
Xanthosoma, Ipomea, Manihot and Maranta cf. arundinacea
(Aceituno et al., 2013: 27) e a list that can also be seen as an
indication of a previous history of exploration of the local resources
(Ichikawa et al., 2011). In this context, it is interesting to propose,
like Mora and Gnecco do, that at sites such as these foragers
promoted the articial concentration of useful plants across their
territory. This farming-like behavior focused on species that
required little planting or tending (Mora and Gnecco, 2003: 282).
However, evidence in support of this is difcult to nd. The presence of pioneer species like Plantago and Trema in a context of

Please cite this article in press as: Borrero, L.A., Moving: Hunter-gatherers and the cultural geography of South America, Quaternary
International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.011

L.A. Borrero / Quaternary International xxx (2014) 1e8

mature primary forest indicates human disturbance but, as


admitted by Gnecco, I cannot say whether this open space was
naturally or humanly created (Gnecco, 2003a: 14). There are also
other evidences of early Holocene habitat transformation at the
tropical forests of northwest Colombia (Gnecco and Aceituno,
2006: 93). Since the presence of humans is normally associated
with disturbance (Odling-Smee et al., 2003), the evidences just
presented are not necessary an indication of particularly complex
interactions with the environment.
It is also argued that the presence of Virola in the pollen sample,
which today is allopatric to the rest of the plants recorded at San
Isidro, may have been transported from their original habitat
(Gnecco, 2003a: 14). This is an expected situation for the PleistoceneeHolocene Transition times under the model of coevolutionary equilibrium (Graham and Lundelius, 1984), that predicts the
existence of non-analog environments for the Late Pleistocene (see
Gnecco, 2003b: 69).
The situation is slightly different for the Amazon basin. Beyond
eating plants and insects, humans hunted a variety of animals
including monkeys, peccaries, tapirs, and others and they are
recorded in the archaeological record (Roosevelt et al., 1996; Politis,
1996a, 2001). These were all new species for foragers coming from
non-forested habitats. Again, the ethology and distribution of these
animals was to be learned in order to make adequate subsistence
choices. Closer to the Atlantic coast, at Lagoa Santa, studies of the
oral health of early Holocene human populations showed that
during the early Holocene the diet was probably based on wild
tubers and fruits (Da Gloria and Larsen, 2014). That early huntergatherers were advanced in the process of managing plant resources at the beginning of the Holocene is an interesting possibility, but it is not yet clear that they were the rst explorers of
those regions. Recent studies suggest that there is a strong possibility that the rst hunter-gatherers in South America already had
bottle gourds, probably used as containers (Erikson et al., 2005;
Piperno, 2011a) and possibly other economically important plants
(Piperno, 2006). Moreover, it is clear that the process of humanizing
the South American environments began with the rst explorers
(Gnecco and Aceituno, 2006: 103), but it probably took some time
to signicantly transform them, before some demographic success
was achieved. Importantly, in terms of the anthropic transformation of forested patches some knowledge is required to select
the locations which are attractive enough to be settled and transformed. It probably took much travel to nd the areas where
burning of the forest is productive, returning was an interesting
option, and human installation was desirable. In other words, the
geography needs to be known before its systematic transformation
fully starts.
What all this evidence clearly indicates, nding also strong
support in the ethnoarchaeological work of Politis among the
Nukak of Colombia, who unintentionally create patches of edible
plants on abandoned camps (Politis and Gamble, 1994; Politis,
1996a, 1996b, 2007), is that the exploitation of the forest environments started very early (Roosevelt et al., 1996; Oliver, 2001; Politis,
2001). Accordingly, there is a good basis to presume that transformations of the forest are also early. This is independent of the
fact that patches of edible plants can also be created in absence of
human occupation (Crdenas and Politis, 2000: 87e88). The mere
presence of humans triggers habitat transformations (Lyman, 1995;
Odling-Smee et al., 2003). The early presence of arrowroot (Maranta sp.) (Piperno, 1995), bottle gourd (Erikson et al., 2005), rhyzomes of Calathea allouia (Stothert et al., 2003; Piperno, 2009) and
other plants in different places of South America (Piperno and
Pearsall, 1998) is central in this discussion. Recently acquired
knowledge about the human ways of exploiting forest resources
derived from evolutionary ecology studies supports a discussion of

the antiquity of these adaptations (Gragson, 1993). Also, the evidence for res in the Amazon basin during much of the Holocene
(Saldarriaga and Clark, 1986; Piperno, 1995), and for processes of
deforestation associated with them (Bray, 1995), suggests that human management existed since early times (Stahl, 1996: 114). All
these processes, however, must be adequately documented in
relation with specic archaeological populations, acknowledging
the existence of a previous process of humans entering a new
environment with new resources, getting used to them, and nally
learning the tactics associated with their management. David Rindos coevolutionary theory might be relevant here, as it requires
substantial time for the establishment of coevolutionary relationships between humans and plants (Rindos, 1984; Gnecco, 2000:
130; Gnecco and Aceituno, 2006: 92). However, Piperno does not
believe that the protracted mutualism involved in the theory of
Rindos can be defended, and instead she sustains that long periods
of experimentation with a fairly large and diverse set of species,
especially those with similar life history and nutritional qualities,
would not occur before the establishment of productive farming
systems (Piperno, 2006: 160). Whatever the outcome of these
alternative positions, what is needed is a better knowledge of the
point at which human populations display what Smith (2001) calls
low-level food production systems. In order to achieve this we will
also need well preserved faunal data from sites in forest contexts
(see Piperno, 2006). However, what the existent archaeological
record shows is slowly unwinding reciprocal plant/human interactions (Piperno, 2011a: S467).
Summing up, it is now generally accepted that the ranking of the
forests as habitat for hunter-gatherers is not necessarily low (Politis
and Gamble, 1994; Denham et al., 2009). Moreover, Piperno
emphasized that the single most important factor driving subsistence changes after the close of the Pleistocene probably was the
dramatic decline in foraging return rates associated with the
demise of glacial-period resources and expansion of forests into
regions where open land vegetation had prevailed during glacial
times (Piperno, 2006: 152), which clearly offers an environmental
context under which management of plants was to be expected.
Taking a global point of view on hunter-gatherers living in forests, it
is possible to say that they were able to rapidly explore and take
advantage of local forest resources (Mercader, 2003: 17), in other
words to live there. However, by taking a closer look it becomes
clear that probably it took several generations of people to adapt to
the tropical forests of Colombia, Venezuela, and Brazil. Research in
the Northwest of South America, plus a series of studies based on
phytoliths are leading this archaeological quest. It is clear that a
process of plant resource management was identied at the PleistoceneeHolocene Transition in South America (Stahl, 1996; Gnecco,
2000; Piperno and Stothert, 2003; Dillehay et al., 2007; Piperno,
2011b) and I believe that the process will be found to be not only
more complex, but also older.
3.2. Non-utilitarian items and exchange
A few places in South America display early evidence of nonutilitarian artifacts. It is not yet clear if these are associated with
an exploratory stage or if they signal the time when effective
colonization was taking place. The examples recorded below might
represent both situations. The examples include the early evidences
of the use of ochre in the Pampas, Argentina (Scalise and Prado,
2006), a bone artifact with incisions at Cueva del Medio, Chile
(Nami, 1994: 159), and a mastodon tusk with geometric designs at
Taguatagua 2, Chile (Nuez et al., 1994). Also, a possible pendant on
a Glossotherium osteoderm at Santa Elina, Brazil (Vilhena Vialou,
1997-1998) and three perforated Mylodon osteoderms at Cueva
de los Chingues, Chile (Martin, 2013) must be considered. It is not

Please cite this article in press as: Borrero, L.A., Moving: Hunter-gatherers and the cultural geography of South America, Quaternary
International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.011

L.A. Borrero / Quaternary International xxx (2014) 1e8

yet clear to what point the selection of sites to live was associated
with non-utilitarian considerations, but the available evidence
establish that other concerns must be also taken into account. This
information is interesting because at some point the so-called
non-utilitarian mobility (Whallon, 2006) becomes important as
a result of proto-exchange webs. The degree to which these systems
developed into full exchange systems is open to question. The evidence to discuss this in depth is rarely published. In the few cases
in which it is available, it points toward a low level of interaction
between distant populations. A rarely considered alternative is that
prehistoric populations were scattered and not necessarily very
interconnected. It is becoming more important to discuss interactions during early times on the basis of specic archaeological
markers, and I will present two examples. One example is provided
by evidence recovered at site QJ-280 on south coastal Peru, dated
between about 11,100 and 10,000 BP. These include tools and debris
on petried wood from a source located at least 20 km in the
interior, an obsidian bifacial tool and debris from the Alca source,
located between 2720 and 5165 masl, and seeds of Opuntia cf. cusindica from environments above 2400 masl (Sandweiss et al., 1998;
Sandweiss and Rademaker, 2011: 284e286). Even if the mode of
acquisition is not clear, the evidence clearly show the interaction
between the highlands and the Peruvian coast during the PleistoceneeHolocene Transition. These results indicate a detailed
knowledge of resources that were available on a variety of environments at different altitudes above the ocean.
Another example is offered by the archaeology of the Ro de La
Plata basin. In this case stone tool assemblages including Fell Cave
projectile points dated between 11,000 and 10,000 BP were found
in the Argentine pampas (Flegenheimer, 1986) and at Urupez,
Uruguay dated between 10,600 and 11,600 BP (Meneghin, 2004;
Nami, 2007, 2013). Those points were also found on surface contexts in both Argentina and Uruguay (Castieira et al., 2011). The
same raw material was used for some tools, including Fell Cave
projectile points, at both sides of the Rio de la Plata, and according
to Flegenheimer et al. (2003) it was collected in Uruguay. Then,
interaction across what is today the Ro de La Plata basin occurred
at such an early time, only that then it was a small river, known as
Paleo-Paran (Bracco et al., 2011). Also, the circulation during the
PleistoceneeHolocene Transition of translucid rocks used for projectile points over distances 140e170 km was recorded (Surez,
2011: 202). This panorama indicates that a detailed knowledge of
the regional environment was probably in place ca. 10,000 BP
(Flegenheimer et al., 2003: 61). Both examples, suggest that the
process of exploration of those sectors of the Pacic and Atlantic
coast respectively were known for quite some time before these
well recorded interactions took place.
3.3. Empty lands
Other evidence of the long processes involved in colonization is
the existence of lands which were not used at all, or only slightly
used during the Holocene. We are not talking about places that lack
systematic research, such as parts of the Caribbean lowlands of
Colombia (Aceituno et al., 2013). Instead, we are referring to areas
that in spite of those efforts are not characterized by an abundant
archaeological record. We must never forget that South America
probably was never fully saturated with people, an important
property that our models must still recognize. Cases like those of
southern Patagonia are good examples. In Santa Cruz, Argentina
there is a large area almost devoid of archaeological materials
located between two nodes of intensive prehistoric occupation
(Borrero and Charlin, 2010). The limited evidence recovered in that
area can be explained as the result of logistical use from one of
those nodes, as a transit zone, or even as a buffer. The main point is

that the area was probably uninhabited most of the time. On the
other hand, work by Mndez et al. (2013) in Aysn, Chile noted the
extremely low frequencies of archaeological remains in the Simpson basin. They entertained the idea of an area demarcating a limit
between populations, but recognized that the evidence is insufcient to discuss it. The implications of very low population densities
are clear. In both examples, archaeologists were at odds to explain
the absence of an archaeological signal. In the end, it probably
marks the existence of very few people with too much available
land. As a result, it needs to be accepted that there are many places
which were populated very late during the Holocene, such as many
dead-end valleys near the Patagonian Cordillera (Borrero, 2004;
Espinosa et al., 2009). In this context we are reminded of the
Nukak conceptualization of space, in which there are places which
are named but were not effectively occupied (Politis et al., 2003:18;
Politis, 2007). In the case of some of the unoccupied lands by the
Nukak, Politis notes that, It is not clear if these unoccupied areas
are the product of recent demographic decline or are simply the
consequence of the traditional Nukak mode of land occupation
(Politis, 2006: 41). Some are places which the Nukak have never or
seldom actually visited (Politis, 2006: 26). It is clear that the
conceptualization of space which is not personally known exists
among hunter-gatherers and we have no major reasons to think
that things were too different at the end of the Pleistocene. On the
contrary, all the available evidence for early settlers of South
America indicates very low demographies. In other words, it is
suggested here that the cultural geography of the early inhabitants
of South America included extensive unoccupied lands, which were
rarely visited, and that only with the passing of time was some
continuity in the distribution of settlement achieved.
4. Conclusions
The process of the peopling of South America was probably slow
and complex. Very little is known of the early stages of appropriation of the land, and adequate methodologies to recognize them
should be rened. Sites attributable to an exploration stage are
elusive, but not unknown, as the examples from Tierra del Fuego
(Piana et al., 2012), Colombia (Aceituno-Bocanegra and Castillo
Espita, 2005), or the Andean mountains (Gil et al., 2011) show. It
was perhaps noted that I have not relied exclusively on Late Pleistocene information or examples to discuss the peopling of South
America. This results from the conviction that it is only by using the
full archaeological record that we are going to understand this
process. Not only there are many places which were for the rst
time visited by humans during the Holocene, but also some that
were visited in earlier times were abandoned after that initial
occupation and perhaps forgotten. This is a condition that opens
the possibility of successive instances of colonization of the same
lands (Franco, 2004).
Acknowledgements
I want to thank the editors of this volume for their invitation to
contribute, and to Cecilia Pallo for her help with the map.
References
Aceituno-Bocanegra, F.J., Castillo Espita, N., 2005. Mobility strategies in Colombias
middle mountain range between the early and middle Holocene. Before
Farming 2005 (2), 1e17 article 2.
Aceituno, F.J., Loaiza, N., Delgado-Burbano, M.E., Barrientos, G., 2013. The initial
human settlement of Northwest South America during the Pleistocene/Holocene transition: synthesis and perspectives. Quaternary International 301, 23e
33.

Please cite this article in press as: Borrero, L.A., Moving: Hunter-gatherers and the cultural geography of South America, Quaternary
International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.011

L.A. Borrero / Quaternary International xxx (2014) 1e8


Anderson, D.G., Gillam, J.C., 2000. Paleoindian colonization of the Americas: implications from an examination of physiography, demography, and artifacts
distribution. American Antiquity 65, 43e66.
Arthur, W.B., 2009. The Nature of Technology. Free Press, New York.
Aschero, C.A., Martinez, J., 2001. Tcnicas de caza en Antofagasta de la Sierra, Puna
Meridional Argentina. Relaciones 26, 215e241.
Beaton, J., 1991. Colonizing continents: some problems from Australia and the
Americas. In: Dillehay, T.D., Meltzer, D.J. (Eds.), The First Americans: Search and
Research. CRC Press, Boca Raton, pp. 209e230.
Belovsky, G.E., 1987. Hunter-gatherer foraging: a linear programming approach.
Journal of Anthropological Archaeology 6, 29e76.
Berger, J., Swenson, J.E., Persson, I.-L., 2001. Recolonizing carnivores and Nave prey:
conservation lessons from pleistocene extinctions. Science 291, 1036e1039.
Binford, L.R., 1991. When the going gets tough, the tough get going: Nunamiut local
groups, camping patterns and economic organisation. In: Gamble, C.,
Boisnier, W.A. (Eds.), Ethnoarchaeological Approaches to Mobile Campsites,
International Monographs in Prehistory, Ann Arbor, pp. 25e138.
Borrero, L.A., 1989. Spatial heterogeneity in Fuego-Patagonia. In: Shennan, S.J. (Ed.),
Archaeological Approaches to Cultural Identity. Unwin Hyman, London,
pp. 258e266.
Borrero, L.A., 1989e1990. Evolucin cultural divergente en la Patagonia austral.
Anales del Instituto de la Patagonia 19, 133e140.
Borrero, L.A., 1994e95. Arqueologa de la Patagonia. Palimpsesto. Revista de
Arqueologa 4, 9e69.
Borrero, L.A., 2001. Regional taphonomy. Background noise and the integrity of the
archaeological record. In: Kuznar, L.A. (Ed.), Ethnoarchaeology of Andean South
America. Contributions to Archaeological Method and Theory, International
Monographs in Prehistory, Ann Arbor, pp. 243e254.
Borrero, L.A., 2004. The archaeozoology of andean Dead Ends in Patagonia: living
near the Continental Ice Cap. In: Mondini, M., Muoz, S., Wickler, S. (Eds.),
Colonisation, Migration, and Marginal Areas. A Zooarchaeological Approach.
Oxbow Books, Oxford, pp. 55e61.
Borrero, L.A., Charlin, J., 2010. Arqueologa del Campo Volcnico Pali Aike,
Argentina. In: Borrero, L.A., Charlin, J. (Eds.), Arqueologa de Pali Aike y Cabo
Vrgenes (Santa Cruz, Argentina). CONICET-IMHICIHU, Buenos Aires, pp. 9e30.
Borrero, L.A., Borrazzo, K., 2013. Oportunismo, exaptaciones y cambio en arqueologa. MS.
Borrero, L.A., Prevosti, F.J., Martin, F.M., 2013. Ranked habitats and the process of
human colonization of South America. Quaternary International 305, 1e4.
Bracco, R., Garca, F., Inda, H., del Puerto, L., Castieira, C., Canario, D., 2011. Niveles
relativos del mar durante el Pleistoceno nal-Holoceno en la Costa de Uruguay.
In: Garca, F. (Ed.), Holoceno en la zona costera de Uruguay. CSIC, Montevideo,
pp. 65e91.
Bray, W., 1995. Searching for environmental stress: climatic and anthropogenic
inuences on the landscape of Colombia. In: Stahl, P.W. (Ed.), Archaeology in
the Lowland American Tropics. Cambridge University Press, Cambridge, pp. 96e
112.
Butzer, K., 1988. A marginality model to explain major spatial and temporal gaps
in the old and new world pleistocene settlement records. Geoarchaeology 3 (3),
193e203.
Crdenas, D., Politis, G., 2000. Territorio, movilidad, etnobotnica y manejo del
bosque de los Nukak orientales. In: Amazonia colombiana. Estudios Antropolgicos 3. Universidad de los Andes, Bogot.
Castieira, C., Cardillo, M., Charlin, J., Baeza, J., 2011. Anlisis de morfometra geomtrica en puntas Cola de Pescado del Uruguay. Latin American Antiquity 22,
335e358.
Chatters, J.C., 2010. Patterns of death and the peopling of the Americas. In: Jimnez
Lpez, J.C., Serrano Snchez, C., Gonzlez Gonzlez, A., Aguilar Orellano, F.J.
(Eds.), III Simposio Internacional El hombre temprano en Amrica. UNAM,
Mxico, pp. 53e74.
Civalero, M.T., Franco, N.V., 2003. Early human occupations in western Santa Cruz
Province, Southernmost South America. Quaternary International 109-110, 77e86.
Clapperton, C., 1993. The Quaternary Geology and Geomorphology of South
America. Elsevier, Amsterdam.
Da Gloria, P., Larsen, C.S., 2014. Oral health of the Paleoamericans of Lagoa Santa,
Central Brazil. American Journal of Physical Anthropology.
Delgado-Burbano, M.E., 2012. Dental and cranio-facial diversity in the Northern
Andes, and the early peopling of South America. In: Miotti, L.L., Salemme, M.,
Flegenheimer, N., Goebel, T. (Eds.), Southbound. Late Pleistocene Peopling of
Latin America Center for the Study of the First Americans. Texas A & M University, pp. 33e37.
Denham, T., Fullagar, R., Head, L., 2009. Plant exploitation on Sahul: from colonisation to the emergence of regional specialisation during the Holocene. Quaternary International 202, 29e40.
Dillehay, T.D., 2000. The Settlement of the Americas. A New Prehistory. Basic Books,
New York.
Dillehay, T.D., Rossen, J., Andres, T.C., Williams, D.E., 2007. Preceramic adoption of
peanut, squash, and cotton in northern Peru. Science 316, 1890e1893.
Dixon, J.E., 2001. Human colonization of the Americas: timing, technology and
process. Quaternary Science Reviews 20, 277e299.
Erikson, D.L., Smith, B.D., Clarke, A.C., Sandweiss, D.H., Tuross, N., 2005. An Asian
origin for a 10,000-year-old domesticated plant in the Americas. Proceedings of
the National Academy of Sciences 102 (51), 18315e18320.
Erlandson, J.M., 2001. The archaeology of aquatic adaptation: paradigms for a new
millennium. Journal of Archaeological Research 9, 287e350.

Espinosa, S.L., Belardi, J.B., Snico, A., 2009. Cun al oeste? Arqueologa del istmo
de la pennsula Maip (lago San Martn, Provincia de Santa Cruz) en su contexto
regional. Arqueologa 15, 187e207.
Flegenheimer, N., 1986. Excavaciones en el sitio 3 de la localidad Cerro La China.
Relaciones de la Sociedad Argentina de Antropologa 17, 7e28.
Flegenheimer, N., Bayn, C., Valente, M., Baeza, J., Femenias, J., 2003. Long distance
tool stone transport in the argentine pampas. In: Miotti, L.A., Salemme, M.C.
(Eds.), South America: Long and Winding Roads for the First Americans at the
Pleistocene/Holocene Transition. Quaternary International 109-110, 49e64.
Franco, N.V., 2003. Es possible diferenciar los conjuntos lticos atribudos a la
exploracin de un espacio de los correspondientes a otras etapas del poblamiento? El caso del extremo sur de Patagonia. Werken 3, 119e132.
Franco, N.V., 2004. LA organizacin tecnolgica y el uso de escalas espaciales
amplias. El caso del sur y oeste de lago Argentino. In: Acosta, A., Loponte, D.,
Ramos, M. (Eds.), Temas de Arqueologa, Anlisis Ltico. Universidad Nacional de
Lujn, Buenos Aires, pp. 101e144.
Gamble, C., 1994. Timewalkers. Allen Lane, London.
Gil, A., Neme, G., Otaola, C., Garca, A., 2011. Registro arqueofaunstico en los Andes
meridionales entre 11,000 y 5000 aos A.P.: Evidencias en Agua de la Cueva e
Sector Sur (Mendoza, Argentina). Latin American Antiquity 22 (4), 595e617.
Gnecco, C., 2000. Ocupacin temprana de bosques tropicales de montaa. Editorial
Universidad del Cauca, Popayn.
Gnecco, C., 2003a. Against ecological reductionism: late pleistocene huntergatherers in the tropical forests of northern South America. In: Miotti, L.A.,
Salemme, M.C. (Eds.), South America: Long and Winding Roads for the First
Americans at the Pleistocene/Holocene Transition. Quaternary International
109-110, 13e21.
Gnecco, C., 2003b. Contra el reduccionismo ecolgico en la arqueologa de
cazadores-recolectores tropicales. Maguar 17, 65e82.
Gnecco, C., Aceituno, J., 2006. Early humanized landscapes in northern South
America. In: Morrow, J., Gnecco, C. (Eds.), Paleoindian Archaeology. A Hemispheric Perspective. University Press of Florida, Gainesville, pp. 86e104.
Gnecco, C., Mora, S., 1997. Late Pleistocene/early Holocene tropical forest occupations at San Isidro and Pea Roja, Colombia. Antiquity 71, 683e690.
Gragson, T., 1993. Human foraging in lowland South America: pattern and process
of resource procurement. Research in Economic Anthropology 14, 107e138.
Graham, D., Lundelius, E., 1984. Coevolutionary disequilibrium and Pleistocene
extinctions. In: Martin, P.A., Klein, R. (Eds.), Quaternary Extinctions: a Prehistoric Revolution. University of Arizona Press, Tucson, pp. 223e249.
Guichn, R., 1994. Antropologa fsica de Tierra del Fuego: Caracterizacin biolgica
de las poblaciones prehispnicas (Tesis de Doctorado). Universidad de Buenos
Aires.
Hofman, J., 2003. Tethered to stone sor freedom to move: folsom biface technology i
regional perspective. In: Soress, M., Dibble, H. (Eds.), Multiple Approaches to the
Study of Bifacial Technologies. University of Pennsylvania Museum of Archaeology and Anthropology Press, Philadelphia, pp. 229e250.
Ichikawa, M., Hattori, S., Yasuoka, H., 2011. Environmental knowledge among
Central African hunter-gatherers: types of knowledge and intra-cultural variations. In: Whallon, R., Lovis, W.A., Hitchcock, R. (Eds.), Information and its Role
in Hunter/Gatherer Bands. Cotsen Institute of Archaeology, Los Angeles,
pp. 117e132.
Kelly, R.L., 1988. The three sides of a biface. American Antiquity 53, 717e734.
Kelly, R.L., 1999. Hunter-gatherer foraging and colonization of the western hemisphere. Anthropologie 37 (1), 143e153.
Kelly, R.L., 2003a. Maybe we do know when people rst came to North America;
and what does it mean if we do? Quaternary International 109-110, 133e145.
Kelly, R.L., 2003b. Colonization of new land by hunter-gatherers: expectations and
implications based on ethnographic data. In: Rockman, M., Steele, J. (Eds.),
Colonization of Unfamiliar Landscapes: the Archaeology of Adaptation. Routledge, London, pp. 44e57.
Kelly, R.L., Todd, L.C., 1988. Coming into the country: early Paleoindian hunting and
mobility. American Antiquity 53, 231e244.
Lpez-Castao, C.E., Cano-Echeverra, M.C., 2012. En torno a los primeros poblamientos en el noroccidente de Sudamrica: Acercamientos des el valle interandino del Magdalena, Colombia. Boletn de Arqueologa PUCP 15, 43e79.
Lyman, R.L., 1995. White Goats, White Lies. University of Utah Press, Salt Lake City.
Martin, F.M., 2013. Tafonoma de la Transicin Pleistoceno-Holoceno en FuegoPatagonia. Interaccin entre humanos y carnvoros y su importancia como
agentes en la formacin del registro fsil. Ediciones de la Universidad de
Magallanes, Punta Arenas.
Martino, L.A., Osella, A., Dorso, C., Lanata, J.L., 2007. Fisher equation for anisotropic
diffusion: simulating South American human dispersals. Physical Review E 76
(031923), 1e6.
Mary-Rousselire, G., 2008 [1980]. Qitdlarssuaq. Lhistoire Dne Migration Polaire.
ditions Poulsen, Paris.
McGhee, R., 1997. Ancient People of the Arctic. UBC Press, Vancouver.
Meltzer, D., 2009. First Peoples in a New World. Colonizing Ice Age America. University of California Press, Berkeley.
Mndez, C., Reyes, O., Trejo, V., Nuevo Delauney, A., 2013. Ocupacin humana de
alto ro Simpson, Aisn (margen occidental de estepa de Patagonia central)
como caso para medir la intensidad de uso de espacios. In: Zangrando, A.F.,
Barberena, R., Gil, A., Neme, G., Giardina, M., Luna, L., Otaola, C., Paulides, S.,
Salgn, L., Tvoli, A. (Eds.), Tendencias terico-metodolgicos y casos de estudio
en la arqueologa de Patagonia, Museo de Historia Natural de San Rafael,
pp. 193e201.

Please cite this article in press as: Borrero, L.A., Moving: Hunter-gatherers and the cultural geography of South America, Quaternary
International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.011

L.A. Borrero / Quaternary International xxx (2014) 1e8

Meneghin, U., 2004. Urupez: Primer Registro radiocarbnico (C-14) para un yacimiento con puntas lticas pisciformes del Uruguay. Fundacin Arqueologa
Uruguaya, Montevideo.
Mercader, J., 2003. Introduction. The Paleolithic settlement of rain forests. In:
Mercader, J. (Ed.), Under the Canopy. The Archaeology of Tropical Rain Forests.
Rutgers University Press, New Brunswick, pp. 1e31.
Miotti, L.L., 2003. Patagonia: a paradox for building images of the rst Americans
during the Pleistocene/Holocene transition. Quaternary International 109-110,
147e174.
Mora, S., Gnecco, C., 2003. Archaeological hunter-gatherers in tropical forests: a
view from Colombia. In: Mercader, J. (Ed.), Under the Canopy. The Archaeology
of Tropical Rain Forests. Rutgers University Press, New Brunswick, pp. 271e290.
Morello, F., 2005. Tecnologa y Mtodos para el desbaste de lascas en el norte de
Tierra Del Fuego: los ncleos del sitio Cabo San Vicente. Magallania 33 (2), 29e
56.
Muscio, H., 2001. Colonizacin humana del NOA y variacin en el consumo de los
recursos: la ecologa de los cazadores-recolectores de la Puna durante la
Transicin Pleistoceno/Holoceno. http://www.naya.org.ar.
Nami, H.G., 1992. Noticia sobre la existencia de tcnica levallois en pennsula
Mitre, extremo sudoriental de Tierra del Fuego. Anales del Instituto de la
Patagonia 21, 73e80.
Nami, H.G., 1994. Resea sobre los avances de la arqueologa nipleistocnica del
extremo sur de Sudamrica. Chungara 26, 145e163.
Nami, H.G., 2007. Research in the middle Negro river basin (Uruguay) and the
Paleoindian occupation of the Southern cone. Current Anthropology 48 (1),
164e174.
Nami, H.G., 2013. Archaelogy, Paleoindian research and lithic technology in the
Middle Negro River, Central Uruguay. Archaeological Discovery 1 (1), 1e22.
Nelson, M.C., 1991. The study of technological organization. In: Schiffer, M.B. (Ed.),
Archaeological Method and Theory 3. University of Arizona Press, Tucson,
pp. 57e100.
Nuez, L., Varela, J., Casamiquela, R., Schiappacasse, V., Niemeyer, H., Villagrn, C.,
1994. Cuenca de Taguatagua en Chile: el ambiente del Pleistoceno y ocupaciones humanas. Revista Chilena de Historia Natural 67, 503e519.
OConnell, J.F., Allen, J., 2012. The restaurant at the end of the Universe: modelling
the colonisation of Sahul. Australian Archaeology 74, 5e17.
Odling-Smee, F.J., Laland, K.N., Feldman, M.W., 2003. Niche Construction. The
Neglected Process in Evolution. Princeton University Press, Princeton.
Oliver, J.R., 2001. The archaeology of forest foraging and agricultural production in
Amazonia. In: McEwan, C., Barreto, C., Neves, E. (Eds.), Unknown Amazon. The
British Museum Press, London, pp. 50e85.
Oliver, J.R., 2008. The archaeology of agriculture in ancient Amazonia. In:
Silverman, H., Isbell, W.H. (Eds.), Handbook of South American Archaeology.
Springer, New York, pp. 185e215.
Orquera, L.A., Piana, E.L., 2009. Sea nomads of the beagle channel in southernmost
South America: over six thousand years of coastal adaptation and stability.
Journal of Island & Coastal Archaeology 4, 61e81.
Piana, E.L., Zangrando, A.F., Orquera, L.A., 2012. Early occupations in Tierra del Fuego
and the evidence from layer S at the Imiwaia I Site (Beagle Channel, Argentina).
In: Miotti, L.L., Salemme, M., Flegenheimer, N., Goebel, T. (Eds.), Southbound.
Late Pleistocene Peopling of Latin America Center for the Study of the First
Americans. Texas A & M University, pp. 171e175.
Piperno, D.R., 1995. Plant microfossils and their application in the new world. In:
Stahl, P.W. (Ed.), Archaeology in the Lowland America Tropics. Cambridge
University Press, Cambridge, pp. 130e153.
Piperno, D.R., 2006. The origins of plant cultivation and domestication in the
neotropics. A behavioral ecological perspective. In: Kennett, D.J.,
Winterhalder, B. (Eds.), Behavioral Ecology and the Transition to Agriculture.
University of California Press, Berkeley, Los Angeles, London, pp. 137e166.
Piperno, D.R., 2009. Identifying crop plants with phytoliths (and starch grains) in
Central and South America: a review and an update of the evidence. Quaternary
International 193, 146e159.
Piperno, D.R., 2011a. The origins of plant cultivation and domestication in the new
world tropics: patterns, process, and new developments. Current Anthropology
52, S453eS470.
Piperno, D.R., 2011b. Northern Peruvian Early and Middle preceramic agriculture in
Central and South American contexts. In: Dillehay, T.D. (Ed.), From Foraging to
Farming in the Andes. Cambridge University Press, Cambridge, pp. 275e284.
Piperno, D.R., Pearsall, D.M., 1998. The Origins of Agriculture in the Lowland Neotropics. Academic Press, San Diego.
Piperno, D.R., Stothert, K., 2003. Phytolith evidence for early Holocene Cucurbita
domestication in southwest Ecuador. Science 299, 1054e1057.
Politis, G., 1996a. Nukak. SINOMI. Centro Amaznico de Investigaciones Cientcas,
Bogot.

Politis, G., 1996b. Moving to produce: nukak mobility and settlement patterns in
Amazonia. World Archaeology 27, 492e511.
Politis, G., 1999. La estructura del debate sobre el poblamiento de Amrica. Boletn
de Arqueologa 14 (2), 25e51.
Politis, G., 2001. Foragers of the Amazon: the last survivors or the rst to succeed?
In: McEwan, C., Barreto, C., Neves, E. (Eds.), Unknown Amazon. The British
Museum Press, London, pp. 26e49.
Politis, G., 2006. The different dimensions of mobility among the Nukak foragers of
the Colombian Amazon. In: Sllet, F., Greaves, R., Yu, P.-L. (Eds.), Archaeology
and Ethnoarchaeology of Mobility. University Press of Florida, Gainesville,
pp. 23e43.
Politis, G., 2007. Nukak. Ethnoarchaeology of an Amazon People. Left Coast Press,
Walnut Creek, CA.
Politis, G., Gamble, C., 1994. La colonizacin de la Floresta tropical Amaznica en
relacin con el poblamiento de Amrica del Sur. Revista del Museo de Historia
Natural de San Rafael 13, 43e47.
Politis, G., Bonomo, M., Prates, L., 2003. Territorio y movilidad entre la costa
atlntica y el interior de la regin pampeana (Argentina). Estudios IberoAmericanos 29 (1), 11e35.
Randall, A.R., Hollenbach, K.D., 2007. Ethnography, analogy, and the reconstruction
of Paleoindian lifeways. In: Walker, R.B., Driskell, B.N. (Eds.), Foragers of the
Terminal Pleistocene in North America. University of Nebraska Press, Lincoln,
pp. 203e225.
Ranere, A.J., Lpez, C.E., 2007. Cultural diversity in the Late Pleistocene/Early Holocene populations of northwest South America and Lower Central America.
International Journal of South American Archaeology 1, 25e31.
Riches, D., 1982. Northern Nomadic Hunter-Gatherer: a Humanistic Approach. Academic Press, London.
Rindos, D., 1984. The Origins of Agriculture: an Evolutionary Perspective. Academic
Press, New York.
Rockman, M., 2003. Knowledge and learning in the archaeology of colonization. In:
Rockman, M., Steele, J. (Eds.), Colonization of Unfamiliar Landscapes: the
Archaeology of Adaptation. Routledge, London, pp. 3e24.
Rogers, A., 1990. Group selection by selective emigration: the effects of migration
and Kin structure. The American Naturalist 135, 398e413.
Roosevelt, A.C., Lima, M., Lopes, C., Michab, M., Mercier, N., Valladas, H., Feathers, J.,
Barnett, W., Imazio, M., Henderson, A., Sliva, J., Chernoff, B., Reese, D.S.,
Holman, J.A., Toth, N., Schick, K., 1996. Paleoindian cave dwellers in the
Amazon: the peopling of the Americas. Science 272, 373e384.
Saldarriaga, J., Clark, D.C., 1986. Holocene res in the Northern Amazon basin.
Quaternary Research 26, 358e366.
Sandweiss, D.H., 2008. Early shing societies in Western South America. In:
Silverman, H., Isbell, W.H. (Eds.), Handbook of South American Archaeology.
Springer, New York, pp. 145e156.
Sandweiss, D.H., McInnis, H., Burger, R.L., Cano, A., Ojeda, B., Paredes, R.,
Sandweiss, M. de C., Glascock, M.D., 1998. Quebrada Jaguay: Early South
American maritime adaptations. Science 281, 1830e1832.
Sandweiss, D.H., Rademaker, K.M., 2011. El poblamiento del sur peruano: costa y
sierra. Boletn de Arqueologa PUCP 15, 275e293.
Scalise, R., Prado, V.D., 2006. Early use of ocher in the Pampean region of Argentina.
Current Research in the Pleistocene 23, 66e68.
Scheinsohn, V., 2003. Hunter-gatherer archaeology in South America. Annual Reviews in Anthropology 32, 339e361.
Smith, B.D., 2001. Low-level food production. Journal of Archaeological Research 9,
1e43.
Speth, J.D., Spielmann, K.A., 1983. Energy source, protein metabolism, and huntergatherer subsistence strategies. Journal of Anthropological Archaeology 2 (1),
1e31.
Stahl, P.W., 1996. Holocene biodiversity: an archaeological perspective from the
Americas. Annual Reviews in Anthropology 25, 105e126.
Stothert, K.E., Piperno, D.R., Andres, T.C., 2003. Terminal Pleistocene/Early Holocene
human adaptation in coastal Ecuador: the Las Vegas evidence. Quaternary International 109-110, 23e43.
Surez, R., 2011. Arqueologa durante la Transicin Pleistoceno-Holoceno en
Uruguay. Componentes paleoindios, organizacin de la tecnologa ltica y
movilidad de los primeros americanos. BAR International Series 2220.
Archaeopress, Oxford.
Vilhena Vialou, A., 1997e1998. Une pendeloque taille dans un os de Glossotherium. In: Encyclopedie Universalis. Universalia, Paris, p. 267.
Whallon, R., 2006. Social networks and information: non-utilitarian mobility
among hunteregatherers. Journal of Anthropological Archaeology 25, 259e270.
White, M.J., Pettitt, P.B., 2011. The British Late Middle Palaeolithic: an interpretative
synthesis of Neanderthal occupation at the northwestern edge of the pleistocene world. Journal of World Prehistory 24, 25e97.

Please cite this article in press as: Borrero, L.A., Moving: Hunter-gatherers and the cultural geography of South America, Quaternary
International (2014), http://dx.doi.org/10.1016/j.quaint.2014.03.011

Anda mungkin juga menyukai