Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Journal of Anthropological Archaeology 49 (2018) 161–172

Contents lists available at ScienceDirect

Journal of Anthropological Archaeology


journal homepage: www.elsevier.com/locate/jaa

The origins and early development of plant food production and farming in T
Colombian tropical forests

Francisco J. Aceitunoa, , Nicolás Loaizaa,b,c
a
Departamento de Antropología, Universidad de Antioquia, Calle 67 No 53-108, AA 1226 Antioquia, Colombia
b
Temple University Department of Anthropology, Philadelphia, PA 19119, USA
c
ISA-INTERCOLOMBIA, Colombia

A R T I C L E I N F O A B S T R A C T

Keywords: This paper concentrates on archaeobotanical evidence for the adoption of plant cultivation in the forests in seven
Colombia regions of Colombia. We present a synthesis and explanation of the evidence we currently have for the process
Food production that involved the adoption of plant cultivation and the development of food production in this area. The use of
Plant cultivation locally available plant foods in these forests is evident by the Pleistocene/Holocene transition. By the Middle
Domestication
Holocene, exogenous plant domesticates were added, including maize, manioc, and possibly common beans. We
further explore available data on other proxies to discuss models to explain the transition from hunting and
gathering to horticulture.

1. Introduction batatas (L) Lam) were domesticated in different regions (e.g. Brücher,
1989; Piperno and Pearsall, 1998: 4–6; Iriarte, 2007, 2009; Rouiller
The origin of food production is one of the major landmarks in et al., 2013). Archaeological data recovered over the past three decades
human evolution (Watson, 2001; Winterhalder and Kennett, 2006). have indicated that Colombia is a key region for understanding the
Plants were domesticated in multiple regions around the globe, be- origin and dispersal of plant cultivation and domestication, as well as
tween five and ten, depending on the author (e.g. Harlan, 1971; their cultural evolutionary consequences throughout the Neotropics
Vavilov, 1992: 126–128; Piperno and Pearsall, 1998; Piperno, 2011a,b; (Cardale et al., 1989; Gnecco and Salgado, 1989; Gnecco and Mora,
Smith, 1998; Diamond, 2002; Zeder et al., 2006; Balter, 2007). One of 1997; Morcote et al., 1998, 2014; Gnecco, 2000, 2003; Aceituno and
these regions – northwestern South America (Andean Colombia and Loaiza, 2014; Santos et al., 2015). We present, analyze and discuss
adjacent Ecuador) – is a belt of tropical forests with dissimilar tem- several lines of information that elucidate the range of adaptive stra-
perature and precipitation regimes. Piperno (2011a,b) has proposed tegies adopted by people living in several regions of Colombian humid
that this region was a center for plant domestication in South America. forests. We show how they developed strategies during the early Ho-
Synergisms between socio-cultural, climatic, and environmental pro- locene and the middle Holocene, setting the scene for the transition
cesses at the late Pleistocene/early Holocene boundary and subse- from lifeways combining hunting, fishing, and gathering to others that
quently in the early and mid-Holocene, favored situations that en- increasingly relied upon a small number of domesticated vegetable
hanced the subsistence value of planting groups of vegetable foods, foods.
rather than gathering them off the landscape (Gnecco, 2003; Gnecco Human groups present on the Colombian landscape since at least
and Aceituno, 2004, 2006; Aceituno and Loaiza, 2014, 2015; Loaiza 12,400 ± 160, years BP (Correal, 1986: 117) (all dates are un-
and Aceituno, 2015; Morcote et al., 2014; Piperno, 2011a,b, 2017; calibrated 14C, unless noted otherwise) responded to the strong en-
Dickau et al., 2015; Loaiza and Aceituno, 2015; Piperno, 1989; Santos vironmental changes of this long period by developing plant manage-
et al., 2015). ment practices that modified the regional and local vegetation and
A growing number of taxonomic and genetic studies in different included the use of cultigens whose initial domestication occurred
parts of the lowland and premontane forests in the Neotropics show outside and even well beyond Colombia, such as manioc and maize
that important food plants such as manioc (Manihot esculenta Crantz), (Piperno, 2011a,b; 2017). These management strategies are the step-
arrowroot (Maranta arundinacea L), cocoyam (Xanthosoma sagittifolium pingstones for the origins of food production and signal the beginning
(L) Schott), leren (Calathea allouia Lindl) and sweet potato (Ipomoea of the Archaic, a Holocene cultural period in which the development of


Corresponding author.
E-mail addresses: francisco.aceituno@udea.edu.co (F.J. Aceituno), nloaiza@temple.edu (N. Loaiza).

https://doi.org/10.1016/j.jaa.2017.12.007
Received 17 January 2017; Received in revised form 5 December 2017
0278-4165/ © 2018 Elsevier Inc. All rights reserved.
F.J. Aceituno, N. Loaiza Journal of Anthropological Archaeology 49 (2018) 161–172

plant domestication and cultivation as well as the insertion of horti- of Sauzalito, El Recreo and El Pital are located. Radiocarbon dates from
culture as a principal player in the subsistence economy. these sites are between 9670 ± 100 years BP and 4090 ± 90 years BP
We use the expression food production in the sense of low-level food (Salgado, 1988–1990, Salgado, 1995) (Fig. 2). Stone tools found at
production (Smith, 2001), defined as an economy that is a mixture of them, including milling stones and handstones, and axes/waisted hoes
practices characteristic of hunting and gathering and agricultural so- suggest plant processing (Herrera et al., 1988; Salgado, 1988–1990).
cieties (Winterhalder and Kennett, 2006: 4). Horticulture should be The last-named is considered a diagnostic tool type for the region and
understood as a form of low-level food production that entails small- has been related to the removal of soils for plant cultivation, harvesting
scale planting of species that range from wild to domesticated (Piperno tubers and roots, and the extraction of starchy hearths from palm trees
and Pearsall, 1998: 6–7; Winterhalder and Kennett, 2006: 4). Following (Cardale et al., 1989; Gnecco and Salgado, 1989). From these sites ar-
these authors, domesticates are new varieties or species of plants or chaeologists also recovered charred seeds from palm trees and avocado
animals created through artificial selection (Winterhalder and Kennett, along with phytoliths identified as palms, bamboos and arrowroot
2006: 3). Domestication then is the process that leads to the evolution Maranta sp. (Piperno, 1985; Piperno and Pearsall, 1998: 202).
by artificial selection of altered organisms that display features ad- It can be argued that the Calima River data revolutionized archae-
vantageous for human environments but that are disadvantageous for ological interpretations during the 1980s and 1990s) because they
surviving in the wild, and this last feature is also known as the do- suggested a model for early hunter-gatherers that stressed plants, as
mestication syndrome (Allaby, 2014). Cultivation should be understood opposed a heavy reliance on hunting as in the Sabana de Bogotá
as the “tending of plants, wild or domesticated” (Winterhalder and (Gnecco and Salgado, 1989). This new model also put forward for the
Kennett, 2006: 4). first time the possibility of plant cultivation in early Holocene occu-
pations in Colombia.
2. Archaeobotanical evidence of plant use during the Archaic
period 2.3. Middle Cauca

At the onset of the Holocene, field evidence shows a rise in regional Following the Cauca River Basin to the north, is found the middle
population as new areas were colonized (Aceituno et al., 2013). The Cauca region (Central Cordillera) (Fig. 1). This region has one of the
evidence to be discussed in this paper is located from south to north of most complete archaeological records in Colombia (26 sites excavated
the Andean region of Colombia following the Cauca River Basin in and dated) and a long occupational sequence that dates back to the
Popayan, Calima River Basin, middle Cauca River Basin (middle Pleistocene/Holocene transition to the middle Holocene (Rojas and
Cauca), Medellin/Porce River Basin. Outside the Andes in the Amazon Tabares, 2000; Rodríguez, 2002; Cano, 2004, 2008; Tabares, 2004;
basin the Peña Roja site is located. Finally we will refer to archae- Aceituno and Loaiza, 2007; Dickau et al., 2015). Most of these sites are
ological evidence in the Sabana de Bogota and middle Magdalena River located between 1400 and 1700 masl in a premontane wet forest
Basin (middle Magdalena) (Fig. 1). (Espinal, 1985). Archaic occupations in the region are dated between
10,619 ± 66 and 4180 ± 70 years BP (Dickau et al., 2015) (Fig. 2).
2.1. Popayan The Middle Cauca lithic industry comprises flaked and ground tools.
Flaked tools were widely used, and some derived from a quartz quarry
Popayan is located in the Central Cordillera in the upper Cauca site at El Antojo where thousands of flakes were collected, some of them
River Basin at about 1600 m above sea level (masl) in a Subandean bifacially reduced (INTEGRAL, 1997; Aceituno and Loaiza, 2007:
forest zone (Fig. 1) (Gnecco, 2000: 17). In this setting is located San 77–78). Two bifacial projectile points: one at the site of El Mirador,
Isidro, a site dated between 10,050 ± 100 and 9530 ± 100 years BP dated 9663 ± 83 years BP, and another at the 39 El Recreo Cancha site
(Gnecco, 2000: 48, Gnecco, 2003) (Fig. 2) and characterized by thou- dated between ca. 8500 and 8000 years BP (Herrera et al., 2011; Dickau
sands of flaked chert and some obsidian artifacts There are retouched et al., 2015). Frequent handstones, milling stones, and waisted tools
and unretouched flakes, lanceolate bifaces and projectile preforms. In probable axes/waisted hoes (Fig. 3: 1–3) were manufactured with local
addition to the knapped artifacts there are edge ground cobbles volcanic rocks obtained in streams (Aceituno and Loaiza, 2007; 2014;
(handstones), flat milling stones, cobbles with concave grooves, and a 2015; Loaiza and Aceituno, 2015).
ground stone axe (Gnecco, 2000: 60–62). Thousands of seeds were re- Macrobotanical remains are scarce in middle Cauca sites, and
covered in association with the artifacts. These included avocado therefore archaeobotanical studies have been based mainly on pollen
(Persea cf. americana), basul (Erythrina cf. edulis), Caryocar spp., Virola and starch grain analysis. Even so, the sheer quantity of data underlines
spp., several palm types, highlighting Acrocomia (Piperno and Pearsall, the importance of this region for the understanding the evolution of
1998: 200; Gnecco, 2000: 67–69). Macrobotanical remains of Lagenaria food production in the northern Andes. Tables 1 and 2 show wild and
sp. were also recovered at San Isidro (Gnecco, 2003; Gnecco and domesticated taxa identified in archaeological sites dated from the early
Aceituno, 2006: 93). Starch grains extracted from an edge ground to the middle Holocene. During the early Holocene, microbotanical
cobble were identified as cf. Xanthosoma/Ipomoea and/or Manihot and data on plant use suggest the earliest stages of a form of low-level food
Maranta (cf. arundinacea), as well as non-identified grasses and legumes production that entailed the selection, propagation, and protection of
(Piperno and Pearsall, 1998: 200). some plants in areas close to settlements. On the other hand, the exo-
According to Gnecco (2000, 2003) San Isidro inhabitants were al- genous origin of maize (Fig. 4: d and k), Manihot cf. esculenta (Fig. 4: a,
ready practicing cultivation at the Pleistocene/Holocene transition. In e, m, and p) and Phaseolus cf. vulgaris (Fig. 4: l and n) confirms that
the pollen record the identification of colonizing plants (i.e. Plantago), horticultural practices were well established by the middle Holocene
grasses and shrubs suggest plot preparation close to the site. Based on throughout the Neotropics. The remaining taxa identified are carbo-
the pollen and archaeobotanical records, Gnecco (2003) suggested that hydrate rich tuberous plants, such as Xanthosoma (Fig. 4: f, h, and j), a
San Isidro inhabitants practiced a form of agroecology (Rindos, 1984) widespread genus across northern South America (Piperno and Pearsall,
based on the selection and cultivation of non-local fruit trees and tubers 1998: 165) that is well represented within the pollen record at the El
around 10,000 years BP. Jazmin site between ca. 9000 and 5000 years BP (Jaramillo and Mejía,
2000a; Aceituno and Loaiza, 2007: 84–86). That suggests these tubers
2.2. Calima River Basin were likely a resource used since the first moments of food production.
Ipomoea (Fig. 4: g) – the genus to which the domesticated species I.
North of Popayan in the Calima River Basin (Western Cordillera) batatas (L) belongs – produces tuberous plants and has been identified
(Fig. 1) in a Subandean moist forest life zone and at 1750 masl, the sites in the middle Cauca. Two wild possible ancestors of I. batatas, I. trifida

162
F.J. Aceituno, N. Loaiza Journal of Anthropological Archaeology 49 (2018) 161–172

Fig. 1. Colombian geography and archaeological regions.

and I. triloba, are found in various regions of the Cordillera Central and 2014).
the Colombian Caribbean region (Roullier et al., 2013). Ipomoea cf The pollen record corresponding to the early Holocene in the middle
batatas has been reported at Middle Holocene in the Canaán site as it Cauca region displays plant families typically associated with anthropic
was at the 39 El Recreo Cancha site (Dickau, 2008) (Table 2). alteration, such as Asteraceae, Melastomataceae, Poaceae and Cecropia
The Dioscorea situation (Fig. 4: c and o) is somewhat similar to that spp., among others (Jaramillo and Mejía, 2000a). Nonetheless, the
of cocoyam. Tuber-producing yams have a wide distribution through available data are not conclusive about the origin of this alteration and
the Neotropics (Brücher, 1989: 19). Dioscorea pollen has been identified we cannot assess if it was due to garden plot preparation, incidentally to
at El Jazmin, Campoalegre and Guayabito sites (Jaramillo and Mejía, camp preparation as shown in ethnographic examples (e.g. Politis,
2000a, 2000b; Aceituno, 2002), while starch grains from tools have 1995, 1996, 2007), natural causes or some combination of these sce-
been recovered at La Pochola, El Jazmin and La Selva sites (Aceituno narios.
and Loaiza, 2014). Of this genus only D. trifida was domesticated in the As noted above, plant use in the middle Cauca increases by the mid-
Neotropics in the wide lowland area of northern Brazil, the Guyanas, Holocene. Towards. 7500/7000 years BP the pollen records from La
Suriname and southern Venezuela (Piperno and Pearsall, 1998: 117; Pochola, El Jazmin and Campoalegre display clear evidence of vege-
Piperno, 2011a). To date little is known of the domestication history of tation alteration as suggested by the dramatic increase in the plant fa-
this plant and we cannot assure if the starch grains identified belong to milies Melastomataceae and Asteraceae as well as the high taxa di-
a species ancestral to the domesticated D. trifida. However, it is likely versity displayed in the record. Associated with this increase the starch
that the species used was part of the group of plants selected in the early grain evidence of Manihot cf. esculenta, maize and Phaseolus cf. vulgaris
moments of food production. strongly suggest the consolidation of a form of food production based
With regard to common beans (Fabaceae: Phaseolus spp.) (Fig. 4: b on horticulture.
and i) we have been unable to determine if the archaeological speci- Maize was not only identified through starch grains (Table 1), but
mens are local or exotic (Aceituno and Loaiza, 2014). The early dates also from pollen, both at La Pochola dated 6743 ± 45 years BP
suggest that it could be a local variety that remains unidentified but (Mercado, 2010) and El Jazmin dated between ca. 7000 and 5000 years
that is not ancestral to the any of the domesticated species P. vulgaris BP (Table 2) (Aceituno et al., 2001; Aceituno, 2002). Originally do-
(common bean) and P. lunatus (lima bean) originated in Mexico and mesticated in the Balsas River Basin of Mexico before 7900 years BP
Peru (Chacón, 2009; Chacón et al., 2005), around the same time the (8700 cal. BP) (Matsuoko et al., 2002; Piperno et al., 2007, Piperno
local variety was being used in middle Cauca (Aceituno and Loaiza, et al., 2009). In any case, the incorporation of maize in the middle

163
F.J. Aceituno, N. Loaiza Journal of Anthropological Archaeology 49 (2018) 161–172

Fig. 2. Radiocarbon dates from regions discussed.

Cauca to local practices played an important role in the consolidation of Isendahl, 2011). The starch grains from the tools “mano 514” (older
horticultural strategies increasing the use of crops with higher return than 7590 years BP), “mano 265” (dated 7080 ± 50 years BP) and
rates in gardens near sites (Dickau, 2005; Aceituno and Loaiza, 2007, “mano” 762 (dated between 6903 ± 45 and 6743 ± 45 years BP)
2014). (Table 2) display the distinctive characteristics of manioc (Aceituno and
Manihot spp. is another important taxa identified in the middle Loaiza, 2014). Manihot spp. pollen was recovered from the Guayabito
Cauca region. Manioc was domesticated in Southwest Brazil in the site dated 4180 years BP (Table 2) (Jaramillo and Mejía, 2000b;
transitional zone between the Cerrado environment and the Amazon Aceituno, 2001, 2002; Aceituno et al., 2001) supporting the idea that
Basin (Olsen and Schall, 1999; Arroyo-Kalin, 2010; Clement, 2010; this plant was cultivated locally as early as the middle Holocene.

Fig. 3. Axes/Waisted Hoes: (1) El Jazmin Level 9 (code 433); (2) La Pochola level 16 (code 282); (3) El Jazmin (surface collection).

164
F.J. Aceituno, N. Loaiza Journal of Anthropological Archaeology 49 (2018) 161–172

Table 1
Starch grains recovered in middle Cauca sites.

14
Site Tool C YBP Xanthosoma spp. Fabaceae Phaseolus spp. Ipomoea spp. Zea mays Dioscorea spp. Manihot spp.

Pch Handst 55 8095 ± 55


Pch Handst 79 8095 ± 55 1
Pch Handst 238 9047 ± 45/6903 ± 45 4 1 1 1
Pch Handst 324 6903 ± 45 2
Pch Handst 458 6903 ± 45 7 1 10
Pch Handst 23 6903 ± 45/6743 ± 45 3 1
Pch Base 665 6903 ± 45/6743 ± 45 9 31 18 3
Pch Handst 762 6903 ± 45/6743 ± 45 4 3
Pch Handst 733 6903 ± 45/6743 ± 45 3
Pch Handst 236 6743 ± 45 2
Pch Handst 1012 6743 ± 45/5922 ± 51 1 2 1
Pch Handst 2168 6743 ± 45/5922 ± 51 1 1
Pch Handst 2154 5922 ± 51
Jaz Base 315 > 7590 5
Jaz Base 514 > 7590 2 1 5
Jaz Handst 265 7080 ± 50 1 5 1
Jaz Handst 830 5625 ± 50 1
Jaz Handst 202 5625 ± 50/7528 ± 51 4 1
La Selva Milling stone 39 8712 ± 60 2 1
La Selva Axe/hoe 90 > 8712 ± 60 1 2 5
La Selva Axe/hoe 281 8712 ± 60 1

Pch: Pochola; Jaz: El Jazmín; handst: handstone.

By ca. 6900–6700 years BP at La Pochola site a large number of and gathering (Johnson, 1983; Posey, 1984; Balée, 1989, 2006; Balée
starch grains display the typical characteristics of common beans and Gely, 1989; Andrade, 1993; Politis, 1996, 2007; Rival, 1998, 2006;
(Phaseolus vulgaris) (Table 1) (Fig. 4: l and n). It has been estimated that Balée and Erickson, 2006; Winterhalder and Kennett, 2006: 4).
by this time beans had been domesticated in the centers of origin and
that therefore there is a good chance that the starch grains from these
2.5. Medellin/Porce River Basin
levels belong to the domesticated species. Nonetheless this is hard to
prove because the starch grains of domesticated and wild varieties of
North of middle Cauca, the Medellin/Porce basin in the Central
common beans are similar and therefore hard to discriminate through
Cordillera has produced ten Archaic sites that display important evi-
starch grains (Piperno and Dillehay, 2008).
dence of food production. In the South of this region (Medellin), three
In summary, the data from middle Cauca suggest that plants began
sites, La Morena, La Blanquita, and Casa Blanca, are located between
to be manipulated by humans in the early Holocene setting the foun-
2000 and 2100 masl, in an Andean life zone (Santos, 2010). The other
dations for the development and establishment of horticulture by the
sites are located further North (Porce) in a tropical forest life zone just
middle Holocene in which tuberous plants played a key role (i.e. the
below 1000 masl and named PIIIOI-52, PIIIOI-59, PIIIOI-40, PIIIOP-61,
preparation of small gardens where wild and domesticated species are
PII-021, PII-045 and PII-107. Lithic technology is similar to that of the
cultivated) (Piperno and Pearsall, 1998: 7; Winterhalder and Kennett,
Calima and middle Cauca regions, comprising handstones, milling
2006: 4; Harris, 2007: 23–24). An important feature of horticulture in
stones, axes, hoes, and unifacial tools. Besides these kinds of tools, there
tropical environments is the high diversity of plants that are cultivated
are thousands of flintknapped quartz tools (such as knives and scrapers)
nearby habitation sites as an economic strategy that also entails hunting
that have been associated with hunting and butchering activities

Table 2
Middle Cauca archaeological sites with associated archaeobotanical evidence and dates.

14
Site C YBP Archaeobotanical evidence Reference

Jaz ca. 9000–5000 Xanthosoma (p); Dioscorea (p) Aceituno and Loaiza (2007)
Jaz > 7590 Phaseolus spp. (s); Dioscorea spp. (s); Manihot spp. (s) Aceituno and Loaiza (2014)
Jaz ca. 7000–5000 Zea mays (p) Aceituno and Loaiza (2007)
Jaz 7080 ± 50 Zea mays (s); Manihot cf. esculenta. (s) Aceituno and Loaiza (2014)
Jaz 5625 ± 50 Manihot spp. (s) Aceituno and Loaiza (2014)
Campoalegre ca. 7600 Dioscorea spp. (p) Aceituno (2001)
Pch 8680 ± 55 Dioscorea spp. (s); Phaseolus spp. (s) Aceituno and Loaiza (2014)
Pch 8095 ± 55 Phaseolus spp. (s) Aceituno (2016)
Pch 9047 ± 45/6903 ± 45 Xanthosoma spp. (s), Fabaceae (s); Zea mays (s) Phaseolus spp. (s) Aceituno (2016)
Pch 6903 ± 45 Xanthosoma spp. (s), Fabaceae (s); Zea mays (s) Aceituno (2016)
Pch 6903 ± 45/6743 ± 45 Xanthosoma spp. Phaseolus cf vulgaris (s) Fabaceae (s); Zea mays (s); Dioscorea spp. (s); Manihot cf Aceituno and Lalinde (2011)
esculenta (s)
Pch 6743 ± 45 Zea mays (s, p) Aceituno and Lalinde (2011)
Pch 6743 ± 45/5922 ± 51 Xanthosoma spp. (s); Phaseolus cf vulgaris (s); Fabaceae (s) Aceituno (2016)
Arrayanes 6520 ± 90 Juglans nigra (m) Rodríguez (1997)
La Selva 8712 ± 60 Phaseolus spp. (s); Dioscorea spp. (s) Aceituno and Loaiza (2014)
Canaán ca.5600 Zea mays (s); Manihot cf esculenta (s); Calathea sp. (s); Ipomoea cf batatas (s) Dickau (2008)
Guayabito 4180 ± 70 Zea mays (p); Manihot esculenta (p); Passiflora spp. (p) Aceituno (2002)

p: pollen.
s: starch grain.
m: macrobotanical remain.

165
F.J. Aceituno, N. Loaiza Journal of Anthropological Archaeology 49 (2018) 161–172

Fig. 4. (a) El Jazmín (code 514) milling stone, Manihot cf esculenta; (b) El Jazmín (code 514) Phaseolus spp.; (c) El Jazmín (code 514) Dioscorea spp.; (d) El Jazmín (code 202) Zea mays; (e)
El Jazmín (code 202) B2 level 16 milling stone, Manihot cf esculenta; (f) El Jazmín (code 30) A1 level 14, handstone, Xanthosoma spp.; (g) El Jazmín (code 265) D2 level 18 handstone,
Ipomoea spp.; (h) La Pochola (code 43) A2 level 19 handstone, Xanthosoma spp; i) La Pochola (code 104) Fabaceae; (j) La Pochola (code 665) B5 level 14 milling stone, Xanthosoma spp.;
(k) La Pochola (code 665) Zea mays; (l) La Pochola (code 762) A’5 level 13 handstone, Phaseolus cf vulgaris; (m) La Pochola (code 762) Manihot cf esculenta; (n) La Pochola (code 665)
Phaseolus cf vulgaris; (o) La Selva (code 90) level 10, scraper, Dioscorea spp.; (p) La Selva (code 90) Manihot spp.

according to the shape (Castillo and Aceituno, 2006; Otero and Santos, (Piperno and Pearsall, 1998: 74; Castillo and Aceituno, 2006). The
2006: 74–86). pollen record does not show evidence of anthropogenic alteration as-
As in middle Cauca, the Medellin/Porce region cultural sequence is sociated with plant use.
long (Fig. 2), allowing an in-depth analysis of the adaptive behaviors The middle Holocene (ca. 7700–4000 years BP) pollen record dis-
involved in the development of food production. This region has also plays major changes that suggest variation in the economic strategies
zooarchaeological and human remains, adding two more lines of evi- adapted by Medellin/Porce region settlers. Towards ca. 7700 years BP
dence. The early Holocene record is represented only by the identifi- there is evidence of vegetation alteration associated with the clearing of
cation of Dioscorea spp. pollen at La Morena dated between plots to favor the growth of food plants from the families Solanaceae,
10,060 ± 60 and 9680 ± 60 years BP (Santos et al., 2015). For this Arecaceae, Annonaceae and Passiflorae, among others (Castillo and
same time it has been suggested for the Porce subregion a broad Aceituno, 2006). Associated with the clearings, the record shows the
spectrum economy that used mostly riverine ecosystems since their appearance of exogenous crops such as maize (pollen, phytoliths, and
resources are richer and more predictable than inland resources starch grains) and Manihot spp. (starch grains), along with other plants

166
F.J. Aceituno, N. Loaiza Journal of Anthropological Archaeology 49 (2018) 161–172

Table 3
Medellín/Porce River sites with associated archaeobotanical evidence and dates.

14
Site C YBP Archaeobotanical evidence Reference

PIIIOI-52 7730 ± 170 Zea mays (p, ph, s) Phaseolus trichocarpus (p) Otero and Santos (2006) and Santos et al.
(2015)
PII-021 ca.7500–6500 Aracaeae (ph); Zea mays (s), Poaceae (ph) Castillo and Aceituno (2006)
PIIIOI-40 6890 ± 40 Zea mays (ph) Otero and Santos (2012)
PII-021 ca.6500–5000 Zea mays (p), Manihot spp. (p, s), Smilax spp. (p) Amaranthus spp. (p); Annonaceae Castillo and Aceituno (2006)
(ph)
PII-107 ca. 5000 Zea mays (p, ph), Manihot spp. (p) Castillo and Aceituno (2006)
PIIIOP-61 4650 ± 70 Persea americana (m); Cucurbitaceae (m) Cardona et al. (2007)
PIIIOI-52 4170 ± 40 Zea mays (p, ph, s) Otero and Santos (2006)
PII-045 ca. 4200–3500 Amaranthus spp. (s), Manihot spp. (s), Zea mays (s); Ipomoea spp. (s); Cucurbitaceae Castillo and Aceituno (2006)
(ph)
La Morena 10.060 ± 60–4170 ± 50 Phaseolus sp. (p) Persea (p) Santos (2010)
La Morena 10.060 ± 60 Dioscorea (s, p) Santos et al. (2015)
9680 ± 60
La Morena 7080 ± 40 Zea mays (p, s) Dioscorea (p) Santos et al. (2015)
La Morena 7080 ± 40 Zea mays (s) Santos (2010)
4170 ± 50
Casa Blanca 4810 ± 70 Zea mays (p) Langebaek et al. (2000)

p: pollen.
s: starch grain.
m: macrobotanical remain.
ph: phytolith.

such as Smilax spp. (pollen), Amaranthus spp. (pollen) and Phaseolus 2.6. Amazon River Basin
trichocarpus (pollen) (Table 3), taxa that may have been cultivated
(Castillo and Aceituno, 2006; Otero and Santos, 2006: 420; Santos Outside the Colombian Andes, in the Caquetá River Basin in the
et al., 2015). All this botanical evidence suggests that, similar to the greater Amazon River Basin lies the Peña Roja site at 100 masl in a
middle Cauca, the Medellin/Porce inhabitants developed horticulture. tropical rain forest life zone and dated between 9920 and 8090 years BP
As mentioned above, the Porce subregion provides two other lines (Fig. 1) (Fig. 2) (Cavelier et al., 1995; Gnecco and Mora, 1997; Mora,
of evidence that enhance our understanding of its Archaic occupations: 2003: 102; Mora and Gnecco, 2003, Morcote et al., 2014: 43). The lithic
the human and faunal remains found at the 021 site, dated assemblage comprises unifacial flakes, choppers, drills, handstones,
7500–5500 years BP (Castillo, 1998: 45) and with the appearance of milling stones, hammer stones, hoes and anvils, manufactured on local
pottery ca. 5000 years BP (Castillo and Aceituno, 2006). Women, men raw materials such as chert, quartz and igneous rocks (Cavelier et al.,
and children were buried individually, in couples and/or in small 1995: 31–32). Thousands of charred seeds were recovered from this
groups, covered with soils and rocks (Castillo, 1998: 47). About 14,000 Archaic occupation that included several palm and fruit trees such as
animal bones were recovered accompanying the human burials, in a Anaueria brasiliensis, Astrocaryum aculeatum, Astrocaryum chambira, As-
mammal fauna dominated by paca (Cuniculus paca), Tome’s spiny rat trocaryum jauari, Bactris sp., Euterpe precatoria, Oenocarpus bacaba, Oe-
(Proechimys semispinosus), black agouti (Dasyprocta fuliginosa), Brazilian nocarpus bataua, Oenocarpus minor, Mauritia flexuosa, Parkia multijuga,
porcupine (Coendou prehensilis) and nine-banded armadillo (Dasypus Inga spp., Passiflora quadrangularis, Brosimum guianense, Sacoglottis spp.
novemcinctus), along with small amounts of carnivore, bird and reptile and Caryocar spp. among others (Morcote et al., 1998, Morcote et al.,
bones (Castillo and Aceituno, 2006). 2014). Phytoliths from squash (Cucurbita spp.), bottle gourd (Lagenaria
To sum up, archaeobotanical and zooarchaeological evidence sug- siceraria) and leren were also identified. The first two phytolith-pro-
gest that horticulture was the base of the economy: small scale culti- ducing taxa are exogenous plants that had to be brought to the region
vation complemented by the gathering and hunting of local resources. for cultivation (Piperno and Pearsall, 1998: 204–205).
This is also supported by trace elements and stable isotopic analysis on Starch grain analysis identified Xanthosoma spp. was also recovered
the human remains. The values for Strontium (558 ppm), Manganese from two lithic tools (hoe and handstone) and dated between 8800 and
(919 ppm), Barium (148 ppm), Vanadium (98 ppm), Zinc (371 ppm), 8730 years BP (Morcote et al., 2014: 44). The evidence suggests that
Cooper (9 ppm) and 13C (−25.09 and −24.62) suggest an elevated Peña Roja settlers were practicing selective management of palm and
consumption of green plants, low consumption of nuts and fiber, mild fruit trees, as well as cultivation of other food plants (Cavelier et al.,
to low meat consumption and very low fish, crustacean and mollusk 1995: 36–41; Morcote et al., 1998; Morcote et al., 2014).
consumption (Beasley et al., 2013; Corti et al., 2013). The Ba/Sr cor-
relation index (−0.58) also suggests a terrestrial diet based on plants
with a low representation of aquatic resources (Trancho et al., 1996; 2.7. The Sabana de Bogota and the Middle Magdalena River Basin
Castillo and Aceituno, 2006).
The act of burying the dead in a single place accompanied by of- Two other regions in Colombia that have provided important but
ferings of several animal species and lithic tools can be interpreted as a incompletely described archaeobotanical data for this period, are the
cultural strategy for appropriating the landscape, creating a bond with Sabana de Bogota and the Middle Magdalena river basin (Fig. 1). These
the ancestors that reinforced the rights of a particular group over the two regions have not yet been identified as being as important for
local resources (Brown, 1995; Castillo and Aceituno, 2006). It has been discussions about the origin of food production. Nonetheless, we pro-
suggested that the use of pottery vessels (called Cancana) was mainly pose that both sites are relevant for understanding Archaic adaptations
ceremonial, i e, for festivities and social exchanges between neigh- in New World tropical forests across the Plate Pleistocene/early Holo-
boring groups as a strategy to minimize conflict (Castillo and Aceituno, cene boundary.
2006). The Sabana de Bogota is located in the Cordillera Oriental at 2600
masl with 11 archaeological sites. This region has the longest occupa-
tional sequence in northwest South America and includes megafauna

167
F.J. Aceituno, N. Loaiza Journal of Anthropological Archaeology 49 (2018) 161–172

butchering in the Pleistocene/Holocene transition to food production in alterations (Sthal, 2016). Collectively, the archaeological evidence from
the upper Holocene, dated from ca. 12,500 to 3000 year BP (Fig. 1) Colombia supports the worldwide hypothesis that even though some
(Fig. 2) (Correal and van der Hammen, 1977; Hurt et al., 1977; Correal, plants and animals were used and domesticated since the Pleistocene/
1982, 1986, 1990, 1993; Ardila, 1984; Groot, 1992, 1995). Correal Holocene transition, the development of low-level food production
(1986: 123–125) interprets that faunal remains associated to Abriense economies is a Holocene adaptation (Pringle, 1998; Richerson et al.,
technology (unifacial retouched flakes belong to the edge-trimmed 2001; Bettinger et al., 2009).
tradition) suggests that hunting played an important role in the The progressive increase in archaeological evidence from ca.
economy and was complemented with the gathering of snails, crabs, 10,500 years BP in several regions suggest the territorial movements
tubers and seeds. Recent lithic use-wear analysis supports the idea that and expansion into other areas through the Colombian Andes. This si-
Abriense technology was used on a broad spectrum of resources that tuation is similar to what happens in the archaeological records of
included wood, bone and leather working, as well as tools that had adjacent regions such as Panama and Ecuador (Cooke, 1992, 2005;
bamboo phytoliths on them, while only one was identified as a butch- Piperno and Pearsall, 1998: 187; Piperno et al., 2000; Piperno, 2006,
ering tool (Nieuwenhuis, 2002: 54–55, 61, 65, 67). The presence of a 2011a,b; Dickau et al., 2007; Pagán-Jiménez et al., 2016).
wide variety of faunal remains that include species from lower regions The archaeobotanical component (macrobotanical remains, pollen,
also supports the idea of a broad spectrum economy that not only fo- starch grains and phytoliths) shows that carbohydrate rich tuberous
cused on major game but also on hunting minor species (Correal, 1990: plants are a constant in the record of these regions. The data analyzed
257). The continuous presence of guinea pig (Cavia porcellus) and the not only suggest the use of specific plant resources but goes further; in
focus on this species over other larger mammals such as white tailed Popayan (Gnecco, 2000, 2003) and the Amazon basin (Morcote et al.,
deer (Odocoileus virginianus) by the middle Holocene species has led to 1998, 2014) researchers have interpreted the data as evidence of early
the hypothesis that guinea pig domestication may have taken place on anthropic alterations as a set of strategies of landscape management
the Sabana de Bogota between ca 3600 and 2500 years BP (Correal and and control of natural resources.
Pinto, 1983; Correal, 1986: 125; Delgado, 2016). In general terms, the taxa identified through macro and micro-
As the Holocene advanced, lithic technology transformed in- botanical analysis, along with the evidence of anthropic alteration in
corporating an increasing amount of tools focused on plant processing, some regions during the early Holocene, suggest that early inhabitants
such as handstones, anvils, hammers and milling stones (Ardila, 1984; were practicing small scale cultivation and plant management as a
Correal, 1990: 257; Correal, 1987; Groot, 1992: 6; Groot, 1995: 54). At strategy to increase the carrying capacity of local ecosystems, allowing
the Checua site edge ground cobbles were recovered along with milling for an increased control of local resources to ensure the procurement of
and hammer stones dated ca. 8500 years BP, that suggest root and tuber basic resources. The Peña Roja site is a good example: Cucurbita sp. and
processing activities (Groot, 1995: 54; Groot, 1992: 6). By ca. bottle gourd phytoliths show that forest dwellers were cultivating both
3800 years BP there is macrobotanical evidence of oca (Oxalis tuberosa) crops and were also using several kinds of palms, as the thousands of
and squash (Cucurbita pepo) according to Correal (1990: 260) and by ca. palm seeds recovered inform. Taken as a whole this suggest that people
3300 years BP, avocado, maize and sweet potato (Correal, 1990: 261), were altering the distribution of natural resources in some sort of forest
suggesting that food production practices based on plant cultivation management that may have further implications on mobility and de-
since the onset of the Holocene. The increase of faunal diversity in mographic growth.
several sites along with the evidence from Checua suggest an expansion For Popayan Gnecco defends the idea of early cultivation and
of the resources used that had a minimum presence in Pleistocene le- management based on taxa such as arrowroot, avocado, and others such
vels, such as the plants that increase in frequency as the Holocene ad- as Xanthosoma or Ipomoea (Gnecco, 2000, 2003; Gnecco and Aceituno,
vances. 2004, 2006), along with the presence of allopatric species in the pollen
The Middle Magdalena (middle Magdalena River Basin) is located record that have been interpreted as brought together by anthropogenic
between the Cordillera Central and Oriental, between 100 and 200 masl action. For middle Cauca the alteration evidence are weaker, but the
in a tropical rain forest environment (Fig. 1). It constitutes another taxa identified dating to the early Holocene (e.g. Fabaceae, Xanthosoma,
region that, even though it has yet to have substantive archaeobotanical Dioscorea, Phaseolus) that endure through the middle Holocene, suggest
data on food production, it is relevant because of the time frame of the that selection through cultivation was a common strategy in the region
seven sites, dated between 10,400 ± 60 and 3130 ± 70 years BP (Aceituno and Loaiza, 2014).
(López, 1995, 1999, 2008; Otero and Santos, 2002) (Fig. 2), and also Assuming that all these Holocene adaptations included resource
the geographical setting. The lithic tools are composed of several fish- control and management, we need to ask: which were the environ-
tail bifacial projectile points, plano-convex scrapers, and many other mental and/or cultural pressures that triggered the adaptive responses
tools that have been associated with Paleoindian adaptations (López, that were a steppingstone for the development of food production,
1999: 100). The fact that this technology remains the same into the domestication and installment of agriculture? The climatic changes
upper Holocene, has lead for the argument that it was a broad spectrum brought by the Pleistocene/Holocene transition included increased
adaptation based on the wide diversity of riverine resources such as rainfall that lasted up to the Holocene Climatic Optimum
reptiles, birds, water mammals, fish, as well as inland plants and ani- (7000–5000 years BP) favoring the rapid and widespread expansion of
mals (Otero and Santos, 2002). To date, evidence for plant use is limited humid forests and decline of open habitats, which severely affected
to a few starch grains recovered from tools (Nieuwenhuis, 2002: 88–89) fauna, especially large mammals (Piperno, 2017; Piperno and Pearsall,
with a handful of lithic tools focused on plant processing (Otero and 1998: 106–107; Vivo and Carmignotto, 2004). Punctuated and/or cy-
Santos, 2002). Lithic use-wear analysis suggest that the stone tools were clical events of different intensities such as earthquakes, volcanic ac-
used mainly to process fish, leather, wood, bone, and marginally other tivity, tsunamis, ENSO, anthropic alterations, among others, need to be
plants (Nieuwenhuis, 2002: 106). added to climatic tendencies that affected the distribution and relative
abundance of flora and fauna species during the Holocene (Sthal,
3. Discussion 2016). The lack of regional scale studies that provide information on
this topic prevents us from assessing how these continental scale
All the data presented above suggest low-level food production changes behaved in local scales and how they affected the distribution
economies that entail strategies such as ecosystem alteration, resource of resources in the diverse environments of Colombia.
management and plant cultivation during the Holocene, a geological Optimal Foraging models has been a powerful theoretical tool to
epoch that brought large scale environmental changes affecting plant explain the adaptive chances that lead some human societies towards
and animal communities alike as well as unprecedented anthropic food production (Winterhalder and Kennett, 2006; Piperno, 2011,

168
F.J. Aceituno, N. Loaiza Journal of Anthropological Archaeology 49 (2018) 161–172

Piperno, 2017; Cooding and Bird, 2016; Stiner and Kuhn, 2016). One of that favored cultivation and the domestication process. These areas
its models Diet Breadth states that when return rates fall lower ranked would be the scenarios were new selective pressures would act upon the
resources will be included in the diet, this may include innovations to phenotypic plasticity of plants and in some case developing domes-
make them stable in the long run that allow individuals to maintain ticates (Piperno, 2017).
stable return rates (Cooding and Bird, 2016). Demographic pressures Plant cultivation was developed in times of environmental and cli-
and resource depression have been the most used factors to explain why matic changes as adaptive responses to the new conditions that affected
human societies started to expand their diets at the Pleistocene/Holo- resource distribution and availability. The inclusion of a wider diversity
cene transition as a prelude to the development food production of plants into the diet was a risk minimizing strategy that aimed to
(Piperno, 1989; Smith and Winterhalder, 1992; Piperno and Pearsall, guarantee resource availability in different parts of the year as well as
1998; Richerson et al., 2001; Winterhalder and Kennett, 2006; Piperno, to provide a form of bait to procure faunal resources (e.g. Linares, 1976;
2011a; Bird et al., 2016; Cooding and Bird, 2016; Stiner and Kuhn, Dufour, 1990: 54). The high diversity of plant taxa, along with the
2016). evidence for alteration and the presence of foreign plants reinforce the
Resource depression has been wildly used as the major pressure that idea of risk minimizing and broad spectrum economies.
drove human populations for the adaptations that led to the develop- The appearance of foreign domestic crops (manioc, maize and likely
ment of food production. In this sense megafaunal extinction has been common beans) between ca. 7700 and 6500 years BP, further supports
assumed as the major change that triggered the adoption of lower that cultivation was part of an adaptive strategy developed by popu-
ranked resources i.e. plants and small animals (Bird et al., 2016). The lations in Colombia during the early Holocene, and maintained through
scarce archaeological evidence of megafauna in Colombia does not the middle Holocene. Furthermore, it supports the idea of early dis-
suggest that human groups were specialized megafauna hunters persal of crops and horticulture through the Neotropics, a food pro-
(Guitérrez (2010)). Megafauna hunting evidence has been found in duction strategy that also involved the tenure of a wide variety of plants
Tibito (Sabana de Bogota), were mastodon (Haplomastodon and Cu- including fruit trees and small plants (Piperno and Pearsall, 1998: 7;
vieronius), horse (Equus sp.) and white tailed deer dated to Harris, 2007: 23–24). In spatial terms this would also mean a higher
11,740 ± 110 BP were recovered (Correal, 1982). Other sites with dependency on cultivated plots, which in turn would reduce mobility
megafaunal evidence associated with artifacts are El Totumo (Cordil- and increase territorial control. Nonetheless, this has regional differ-
lera Oriental) and La Pileta (Santander) (Correal, 1993); Toro, between ences: while in the middle Cauca there are no clear differences on site
Calima and Middle Cauca in the Cauca River Basin (Rodriguez, 2002, p. sizes through time, in the Porce region there is a high investment on site
29), and Yumbo, south of Toro also in the Cauca River Basin preparation that involved chipped stone floors for habitation areas as
(Rodriguez, 2002, p. 33). Unfortunately, none of these sites have un- well as a specific place to bury the dead between ca. 7700 and
disputable evidence of human activities nor have they been dated, and 5500 years BP.
stratigraphy has only been reported for the first two. The archaeobotanical data recovered during the last 30 years
It is likely that the shifts in density and distribution of small fauna highlight the role of Colombia in the domestication of plants, since the
during the Pleistocene/Holocene transition had a higher impact on macro region has been proposed as the domestication center for several
human populations than megafaunal extinction. South America went crops including cocoyam, leren, arrowroot and sweet potato (Piperno
from 20 mammalian orders in the Pleistocene to 12 in current times, and Pearsall, 1998: 164–165) within region D1 (Piperno, 2011a). Al-
which means a 40% reduction (Vivo and Carmignotto, 2004). Holocene though, the exact location of where domestication took place remains
mammals are usually small and solitary and because of the expansion of unknown, microbotanical analysis done on archaeological sites have
humid forests in this time frame there is an associated expansion of been able to identify all of the genera that those crops belong to dating
fauna that has arboreal locomotion (Piperno and Pearsall, 1998: 62; back to the early and middle Holocene. Xanthosoma has been identified
Vivo and Carmignotto, 2004). Human groups must have been sensitive in tools in the Amazon Basin at ca. 8800 year BP (Morcote et al., 2014:
to the changes in densities of minor fauna. The absence of studies 44), middle Cauca in several tools dating between ca. 9000 and
dealing with small fauna in Colombia difficult makes it hard to assess 5500 years BP, as well as in the pollen record at El Jazmin between ca.
the real weight of this variable, because paleontological studies have 9000 and 5000 years BP (Jaramillo and Mejía, 2000a; Aceituno and
centered mostly on megafauna. As an example, the archaeological re- Loaiza, 2007: 84–86; Aceituno and Lalinde, 2011; Aceituno, 2016), and
cord from Sabana de Bogotá, despite the extinction of horse and mas- a tentative identification for Popayan dated ca. 9500 years BP sup-
todon, suggests that the area had wide terrestrial and aquatic resources porting this idea (Piperno and Pearsall, 1998: 200). The genus Maranta
that help explaining the continuity in human occupations through the was identified through phytoliths from soils in Calima dated between
Holocene. ca. 9000 and 4500 years BP (Piperno, 1985; Piperno and Pearsall, 1998:
There is evidence on territorial expansion into other areas without 202), and tentatively in phytoliths from a tool residue at Popayan
since the Pleistocene/Holocene transition in Colombia (Aceituno et al., dating ca. 9500 years BP (Piperno and Pearsall, 1998: 200). Calathea
2013), but there is no clear evidence that there was population grouth has been identified in a tool from middle Cauca dated ca. 5600 years BP
enough to cause the demographic pressures required to make an impact (Dickau, 2008), and from phytoliths from soils in the Amazon basin
on resources. The estimation of this variable requires more detailed dating ca. 8800 years BP (Piperno and Pearsall, 1998: 204–205). Ipo-
paleoecological and archaeological information that is unavailable at moea from starch grains in the Middle-Cauca, dated ca. 7000 years BP
the time. and Medellin/Porce region dated between ca. 4200 and 3500 years BP
Rather than a single factor, we suspect that a combination of them (Castillo and Aceituno, 2006). As mentioned previously taxonomical
including megafaunal extinction (where present), redistribution of assessment has only been determined to genus, and in some cases is just
ecological communities (plants and animals), demographic growth, tentative, this is due to the fact that microfossil comparative collections
territorial expansion, among others, were the trigger to develop new need to be improved to reach more precision and that sometimes the
behaviors, such as the inclusion of other resources into the diet. In turn, level of species and assessment of wild vs. domesticated cannot be
this led to the selective pressures over those resources and the alteration reached.
of niches allowing for greater control over some animal and plant po-
pulations. Is in this scenario where plant cultivation emerges as an 4. Conclusions
adaptive strategy with ecological consequences that benefited human
populations, including the increase in carrying capacity, reduction of Archaeobotanical data published over the past 30 years have posi-
resource acquisition costs and increase in predictability. The anthro- tioned Colombia as an independent center for the origin of plant use
pogenic alteration of ecosystems allowed for environmental conditions and domestication (Piperno and Pearsall, 1998: 165; Piperno, 2011;

169
F.J. Aceituno, N. Loaiza Journal of Anthropological Archaeology 49 (2018) 161–172

Piperno et al., 2017). Several species have been suggested as being Aceituno, F.J., Loaiza, N., 2014. Early and Middle Holocene evidence for use of plants and
brought under domestication in this part of the world including ar- cultivation in the the Middle Cauca River Basin, Cordillera Central (Colombia). Quat.
Sci. Rev. 86, 49–62.
rowroot, cocoyam, leren and sweet potato, among others. All this bo- Aceituno, F.J., Loaiza, N., 2015. The role of plants in the early human settlement of
tanical evidence suggests that the food production strategy of the in- Colombia. Quat. Int. 363, 20–27.
habitants of different regions responded to specific local agroecologies. Aceituno, F.J., Loaiza, N., Delgado, E., Barrientos, G., 2013. The initial human settlement
of Colombia during the Pleistocene/Holocene transition: synthesis and perspectives.
Food production in forests in broken terrain would have been more Quat. Int. 301, 23–33.
microzonal than in wide colluvial valleys, volcanic basins or along Aceituno, F.J., Loaiza, N., Jaramillo, A., Vélez, L., 2001. Identificación de plantas ali-
coasts. menticias en el Cauca medio durante el Holoceno temprano y medio. Boletín de
Antropología 15 (32), 51–72.
The data discussed above show that since the Pleistocene/Holocene Andrade, A., 1993. Sistemas agrícolas tradicionales en el medio río Caquetá. In: Correa, F.
transition human groups altered local environments in order to increase (Ed.), La selva humanizada: ecología alternativa en el trópico húmedo colombiano.
human control over natural resources especially useful plants that ICANH, FEN Colombia, CEREC, Bogotá, pp. 63–86.
Allaby, R.G., 2014. Domestication syndrome in plants. In: Smith, C. (Ed.), Encyclopedia of
present a disperse distribution in tropical forests. Following the opti-
Global Archaeology. Springer, New York, pp. 2182–2184.
mization principle, cultural selection favored behaviors that enhanced Ardila, G., 1984. Chia, un Sitio Precerámico en la Sabana de Bogotá. Fundación de
prediction and control and decreased procurement time. The alteration Investigaciones Arqueológicas Nacionales, Banco de la República, Bogotá.
in distribution and diversity of plants may have also served as a way to Arroyo-Kalin, M., 2010. The Amazonian formative: crop domestication and anthro-
pogenic soils. Diverstiy 2, 473–504.
increase the predictability on animal behavior, increasing the chances Balée, W., 1989. The culture of Amazonian forests. Adv. Econ. Bot. 7, 1–121.
of encounter. The result of combining all these strategies is the devel- Balée, W., 2006. The research program of historical ecology. Annu. Rev. Anthropol. 35,
opment of food producing economies. By the middle Holocene, horti- 75–98.
Balée, W., Gely, A., 1989. Managed forest sucession in Amazonia: the Ka’apor Case. Adv.
culture consolidates all of the strategies as the main form of food pro- Econ. Bot. 7, 129–158.
duction that extends to most of the regions discussed in the previous Balée, W., Erickson, C., 2006. Time, complexity and historical ecology. In: Balée, W.,
sections. The development and implementation of horticultural prac- Erickson, C. (Eds.), Time. Columbia University Press, Complexity and Historical
Ecology studies in the Neotropical Lowlands, pp. 1–17.
tices must have altered population growth, mobility, and territoriality Balter, M., 2007. Seeking agriculture’s ancient roots. Science 316, 1830–1835.
in different ways in each region paving the way for the establishment of Beasley, M.M., Martínez, A.M., Simons, D.D., Bartelink, E.J., 2013. Paleodietary analysis
agricultural societies, but those effects are beyond the scope of this of a San Francisco bay area shellmound: stable carbon and nitrogen isotope analysis
of late Holocene humans from the Ellis Landing site (CA-CCO-295). J. Archaeol. Sci.
paper. 40 (4), 2084–2094.
The progressive increase in use, control and management of plants Bettinger, R., Richerson, P., Boyd, R., 2009. Constraints on the development of agri-
is associated to the environmental changes during the Pleistocene/ culture. Curr. Anthropol. 50 (5), 627–631.
Bird, D.W., Bliege, R., Cooding, B.F., 2016. Pyrodiversity and the Anthropocene: the role
Holocene transition that affected resource distribution and structure.
of fire in the broad spectrum revolution. Evol. Anthropol. 25, 105–116.
The increased dependence on plants along with the strategies devel- Brown, J.A., 1995. On mortuary analysis with special reference to the Saxe-Binford re-
oped to manipulate them, are part of the behavioral adaptations that search program. In: Beck, L.A. (Ed.), Regional Approaches to Mortuary Analysis.
human societies have to ensure survival. Cultural adjustments have an Plenum Press, New York, pp. 3–26.
Brücher, H., 1989. Useful Plant of Neotropical Origin and Their Wild Relatives. Springer,
impact in local ecosystems just as the latter have in the former, creating Berlin.
codependency in this new ecological relation leading in some cases to Cano, M.C., 2004. Los primeros habitantes de las cuencas medias de los ríos Otún y
domestication. The arrival of foreign resources towards the middle Consota. In: López, C.E., Cano, M.C. (Eds.), Cambios Ambientales en Perspectiva
Histórica: Ecorregión del Eje Cafetero. Universidad Tecnológica de Pereira y
Holocene adopted into local practices reinforces the idea that native Programa Ambiental GTZ, Pereira, Colombia, pp. 68–91.
adaptations played a fundamental role in the development of food Cano, M.C., 2008. Evidencias Precerámicas en el Municipio de Pereira: Efectos del
production strategies in Colombia. Vulcanismo y Colonización Temprana de los Bosques Ecuatoriales en el Abanico
Fluviovolcánico Pereira-Armenia. In: López, C., Ospina, G. (Eds.), Ecología Histórica:
Interacciones Sociedad Ambiente a Distintas Escalas Socio temporales. Universidad
Acknowledgements Tecnológica de Pereira-Universidad del Cauca-Sociedad Colombiana de Arqueología,
Pereira, pp. 149–168.
Cardale, M., Bray, W., Herrera, L., 1989. Reconstruyendo el pasado en Calima resultados
We want to thank the anonymous reviewers for their valuable
recientes. Boletín del Museo del Oro 24, 3–33.
comments and suggestions. NL wants to thank Wenner-Gren Cardona, L.C., Nieto, L.E., Pino, J.I., 2007. Del Arcaico a la Colonia. Construcción del
Foundation for the Wadsworth International Fellowship and paisaje y cambio social en el Porce Medio. Informe final. Universidad de Antioquia-
Empresas de Medellín, Medellín (unpublished manuscript).
Colciencias for the Beca Francisco Jose de Caldas, both aimed to sup-
Castillo, N., 1998. Antiguos Pobladores del Valle Medio del Río Porce. Empresas Públicas
port his doctoral studies. Both authors want to thank Vicerrectoría de de Medellín, Medellín.
Investigación (Universidad de Antioquia) and the Research Group Castillo, N., Aceituno, F.J., 2006. El bosque domesticado, el bosque cultivado: Un proceso
Medioambiente y Sociedad for their support. milenario en el valle medio del rio Porce en el noroccidente Colombiano. Latin Am.
Antiquity 17, 561–578.
Cavelier, I., Rodríguez, C., Herrera, L., Morcote, G., Mora, S., 1995. No solo de la caza
Appendix A. Supplementary material vive el hombre: Ocupación del bosque amazónico, Holoceno temprano. In: Cavelier,
I., Mora, S. (Eds.), Ámbito y Ocupaciones Tempranas de la América Tropical.
Fundación Erigaie, Instituto Colombiano de Antropología, Bogotá, pp. 27–44.
Supplementary data associated with this article can be found, in the Chacón, M.I., 2009. Darwin y la domesticación de plantas en las Américas: el caso del
online version, at http://dx.doi.org/10.1016/j.jaa.2017.12.007. maíz y el fríjol. Acta Biológica Colombiana 14 (4s), 351–364.
Chacón, M.I., Pickersgill, B., Debouck, D.G., 2005. Domestication patterns in common
bean (Phaseolus vulgaris L.) and the origin of the Mesoamerican and Andean culti-
References vated races. Theor. Appl. Genet. 110, 432–444.
Clement, C., 2010. Origin and domestication of native Amazonian crops. Diversity 2,
Aceituno, F.J., 2001. Ocupaciones Tempranas del Bosque Tropical Subandino en la 72–106.
Cordillera Centro-Occidental de Colombia. Doctoral Dissertation. Universidad Cooding, B.F., Bird, D.W., 2016. Behavioral ecology and the future of archaeological
Complutense de Madrid, Madrid. science. J. Archaeol. Sci. 56, 9–20.
Aceituno, F.J., 2002. Interacciones fitoculturales en el Cauca medio durante el Holoceno Cooke, R.G., 1992. Etapas tempranas de la producción de alimentos vegetales en la baja
temprano y medio. Arqueología del Área Intermedia 4, 89–113. Centroamerica y partes de Colombia (Región histórica Chibcha-Choco). Revista de
Aceituno, F.J., 2016. Nuevos Datos para La Arqueología Temprana del Cauca Medio: La Arqueología Americana 6, 35–70.
Pochola un sitio precerámico en La Cordillera Central De Colombia, Universidad de Cooke, R.G., 2005. Prehistory of native americans on the Central American land bridge:
Antioquia, Medellín. Report to Instituto Colombiano de Antropología e Historia colonization, dispersal and divergence. J. Archeol. Res. 13 (2), 129–187.
(unpublished manuscript). Corti, C., Rampazzi, L., Ravedoni, C., Giussani, B., 2013. On the use of trace elements in
Aceituno, F.J., Lalinde, V., 2011. Residuos de almidones y el uso de plantas durante el ancient necropolis studies: overview and ICP-MS application to the case study of
Holoceno medio en el Cauca medio. Caldasia 33 (1), 1–20. Valvadaro site, Italy. Microchem. J. 110, 614–623.
Aceituno, F.J., Loaiza, N., 2007. Domesticación del Bosque en el Cauca Medio Correal, G., 1982. Restos de megafauna asociados a artefactos en la Sabana de Bogotá.
Colombiano entre el Pleistoceno Final y el Holoceno Medio. British Archaeological Caldasia XIII 64, 487–547.
Reports. International Series 1654. Archaeopress, Oxford. Correal, G., 1986. Apuntes sobre el medio ambiente pleistocénico y el hombre

170
F.J. Aceituno, N. Loaiza Journal of Anthropological Archaeology 49 (2018) 161–172

prehistórico en Colombia. In: Bryan, A.L. (Ed.), New Evidence for the Pleistocene Jaramillo, A., Mejía, JC., 2000a. Análisis palinológico de los yacimientos el Jazmín y
Peopling of the Americas. Center for the Study of the Early Man, University of Main, Guayabito. Departamento de Risaralda, Medellín (unpublished manuscript).
Orono, pp. 115–131. Jaramillo, A., Mejía, J.C., 2000b. Análisis palinológico del yacimiento Campoalegre.
Correal, G., 1990. Evidencias culturales durante el Pleistoceno y Holoceno de Colombia. Departamento de Risaralda, Medellín (unpublished manuscript).
Revista de Arqueología Americana 1, 69–89. Johnson, A., 1983. Machiguenga gardens. In: Hames, R.B., Vickers, W.T. (Eds.), Adaptive
Correal, G., 1993. Nuevas evidencias culturales pleistocénicas y megafauna en Colombia. Responses of Native Amazonians. Academic Press, New York, pp. 29–63.
Boletín de Arqueología 8, 3–12. Langebaek, C., Espinosa, I., Giraldo, S., 2000. Prospección arqueológica del valle de
Correal, G., Pinto, M., 1983. Investigaciones Arqueológicas en el municipio de Zipacón, Aburrá y sus ecosistemas estratégicos. Estudios de cambios sociales en una región del
Cundinamarca. Fundación de Investigaciones Arqueológicas Nacionales, Banco de la Occidente de Colombia. Informe final. Área Metropolitana del Valle de Aburrá-
República, Bogotá. CORANTIOQUIA-STRATA ltda-CESO, Medellín (unpublished manuscript).
Correal, G., van der Hammen, T., 1977. Investigaciones Arqueológicas en los Abrigos Linares, O.F., 1976. “Garden Hunting” in the American tropics. Hum. Ecol. 4 (4),
Rocosos del Tequendama. Biblioteca Banco Popular, Bogotá. 331–349.
Delgado, M.E., 2016. Stable isotope evidence for dietary and cultural change over the Loaiza, N., Aceituno, F.J., 2015. Reflexiones en torno al Arcaico colombiano. Revista
Holocene at the Sabana de Bogotá region, Northern South America. Archaeol. Colombiana de Antropología 51 (2), 121–146.
Anthropol. Sci. http://dx.doi.org/10.1007/s12520-016-0403-3. López, C., 1995. Dispersión de puntas de proyectil bifaciales en la cuenca media del río
Diamond, J., 2002. Evolution, consequences and future of plant and animal domestica- Magdalena. In: Cavelier, I., Mora, S. (Eds.), Ámbito y Ocupaciones Tempranas de la
tion. Nature 418 (6898), 700–707. América Tropical. Fundación Erigaie, Instituto Colombiano de Antropología, Bogotá,
Dickau, R., 2005. Resource use crop dispersals, and the transition to agriculture in pre- pp. 73–82.
historic Panama: evidence from starch grains and macroremains. Doctoral López, C., 1999. Ocupaciones Tempranas en las Tierras Bajas Tropicales del valle Medio
Dissertation. Temple University, Philadelphia. del Río Magdalena Sitio 05-YON-002, Yondó-Antioquia. Fundación de
Dickau, R., 2008. El uso de maíz y cultígenos de raíces en el precerámica de Panamá y Investigaciones Arqueológicas Nacionales, Banco de la República, Bogotá.
Colombia: evidencia de almidones en sitios húmidos subtropicales premontanos. In: López, C., 2008. Landscape Development and the Evidence for Early Human Occupation
López, C., Ospina, G. (Eds.), Ecología Histórica: Interacciones Sociedad Ambiente a in the Inter-Andean Tropical Lowlands of the Magdalena River.
Distintas Escalas Socio temporales. Universidad Tecnológica de Pereira-Universidad Colombia.SyllabaPress, Miami.
del Cauca-Sociedad Colombiana de Arqueología, Pereira, pp. 60–67. Matsuoko, Y., Vigouroux, Y., Goodman, M.M., Sanchez, J., Buckler, E., Doebley, J., 2002.
Dickau, R., Ranere, A.J., Cooke, R.G., 2007. Starch grain evidence for the preceramic A single domestication for maize shown by multilocus microsatellite genotyping.
dispersals of maize and root crops into tropical dry and humid forest of Panama. PNAS 99 (9), 6080–6084.
PNAS 14 (9), 3651–3656. Mercado, J., 2010. Análisis polínico en el yacimiento la Pochola: un contexto precerámico
Dickau, R., Aceituno, F.J., Loaiza, N., López, C.E., Cano, M., Herrera, L., Restrepo, J.C., en el río San Eugenio (Cauca medio, Risaralda –Colombia-). Magister Tesis. Instituto
Ranere, A.J., 2015. Radiocarbon chronology of preceramic occupation in the Middle de Biología, Universidad de Antioquia (unpublished results).
Cauca Valley, Colombia. Quat. Int. 363, 43–54. Mora, S., 2003. Habitantes Tempranos de la Selva Tropical Lluviosa Amazónica un
Dufour, D.L., 1990. Uso de la selva tropical por los indígenas Tukano del Vaupés. In: Estudio de las Dinámicas Humanas y Ambientales. Pittsburgh, Universidad Nacional
Correa, F. (Ed.), La Selva Humanizada: Ecología Alternativa en el Trópico Húmedo de Colombia. Sede Leticia Instituto Amazónico de Investigaciones. IMANI University
colombiano. Instituto Colombiano de Antropología –ICAN, Fondo FEN Colombia, of Pittsburgh, Department of Anthropology, Latin American Archaeology Reports
Fondo Editorial CEREC, pp. 47–62. n. 3.
Espinal, L.S., 1985. Geografía ecológica del Departamento de Antioquia (zonas de vida Mora, S., Gnecco, C., 2003. Archaeological hunter-gatherers in tropical rainforests: A
formaciones vegetales) del Departamento de Antioquia. Revista Facultad Nacional de view from Colombia. In: Mercader, J. (Ed.), Under the Canopy: the Archaeology of
Agronomía 38 (1), 5–106. Tropical Rain Forests. Rutgers University Press, New Brunswick, pp. 271–290.
Gnecco, C., 2000. Ocupación Temprana de Bosques Tropicales de Montaña. Editorial Morcote, R.G., Cabrera, C., Mahecha, D., Franky, C., Cavelier, I., 1998. Las palmas entre
Universidad del Cauca, Popayán. los grupos cazadores recolectores de la Amazonia colombiana. Caldasia 20, 57–74.
Gnecco, C., 2003. Against ecological reductionism: Late Pleistocene hunter-gatherers in Morcote, G., Aceituno, F.J., León, T., 2014. Recolectores del holoceno temprano en la
the tropical forests of northern South America. Quat. Int. 109–110, 13–21. floresta amazónica colombiana. In: Rostain, S. (Ed.), Antes de Orellana. Actas del 3er
Gnecco, C., Aceituno, F.J., 2004. Poblamiento temprano y espacios antropogénicos en el Encuentro Internacional de Arqueología Amazónica. Instituto Francés de Estudios
norte de Sudamérica. Complutum 15, 151–164. Andinos; Facultad Latinoamericana de Ciencias Sociales, Quito, pp. 39–50.
Gnecco, C., Aceituno, F.J., 2006. Early humanized landscapes in Northern South America. Nieuwenhuis, C.J., 2002. Traces on Tropical Tools: a Functional Study of Chert Artifacts
In: Morrow, J.E., Gnecco, C. (Eds.), Paleoindian Archaeology: A Hemispheric from Preceramic Sites in Colombia. PhD Thesis from Leiden University.
Perspective. University Press of Florida, Gainesville, pp. 86–104. Archaeological Studies Leiden University No. 9. Faculty of Archaeology, University of
Gnecco, C., Mora, S., 1997. Late Pleistocene/Early Holocene tropical forest occupations at Leiden, Leiden.
San Isidro and Peña Roja, Colombia. Antiquity 71, 683–690. Olsen, K.M., Schall, B., 1999. Evidence of origin of cassava: phylogeography of Manihot
Gnecco, C., Salgado, H., 1989. Adaptaciones precerámicas en el suroccidente de esculenta. PNAS 96, 5586–5591.
Colombia. Boletín del Museo del Oro 24, 35–55. Otero, H., Santos, G., 2002. Aprovechamiento de recursos y estrategias de movilidad de
Groot, A.M., 1992. Checua: Una Secuencia Cultural entre 8500 y 3000 Años Antes del los grupos cazadores-recolectores holocénicos del valle medio del Magdalena,
Presente. Fundación de Investigaciones Arqueológicas Nacionales, Banco de la Colombia. Boletín de Antropología 16 (33), 100–134.
República, Bogotá. Otero, H., Santos, G., 2006. Las ocupaciones prehispánicas del cañón del río Porce.
Groot, A.M., 1995. Checua: un aporte para el conocimiento de la Sabana de Bogotá. In: Prospección, rescate y monitoreo arqueológico. Proyecto hidroeléctrico Porce III.
Cavelier, I., Mora, S. (Eds.), Ámbito y Ocupaciones Tempranas de la América Obras de infraestructura. Informe final. Universidad de Antioquia-Empresas de
Tropical. Fundación Erigaie, Instituto Colombiano de Antropología, Bogotá, pp. Medellín, Medellín (unpublished manuscript).
45–58. Otero, H., Santos, G., 2012. Porce III proyecto hidroeléctrico estudios de arqueología
Guitérrez, J., 2010. Erase una vez en Colombia: la megafauna suramericana durante el preventiva. Dinámica de cambio en las sociedades prehispánicas de la cuenca baja del
proceso de poblamiento del cono sur. Arqueología del Area Intermedia 8, 11–82. Porce. Empresas Públicas de Medellín, Medellín.
Harlan, J.R., 1971. Agricultural origins: centers and noncenters. Science 174, 468–474. Pagán-Jiménez, J.R., Guachamín-Tello, A.M., Romero-Bastidas, M.E., Constantine-Castro,
Harris, D., 2007. Agriculture, cultivation and domestication exploring the conceptual A.R., 2016. Late ninth millenium B.P use of Zea mays L. at Cubilán area, highland
framework of early food production. In: Denham, T., Iriarte, J., Vrydaghs, L. (Eds.), Ecuador, revealed by ancient starches. Quat. Int. 404, 137–155.
Rethink Agriculture: Archaeological and Ethnoarchaeological Perspectives. One Piperno, D.R., 1985. Phytolith records from prehistoric raised fields in the Calima región,
World Archaeology, Left Coast Press, California, pp. 16–35. Colombia. Pro-Calima 4, 37–40.
Herrera, L., Bray, W., Cardale, M., Botero. P., 1988. Nuevas fechas de radiocarbono para Piperno, D.R., 1989. Non-affluent foragers: resource availability, seasonal shorteges, and
el precerámico en la Cordillera Occidental de Colombia. In: Paper Presented at the the emergence of agriculture in Panamanian tropical forests. In: Harris, D.R.,
46th International Congress of Americanists, Amsterdam. Hillman, G.C. (Eds.), Foraging and Farming: The Evolution of Plant Exploitation. One
Herrera, L., Moreno, M.C., Peña, O., 2011. La Historia muy Antigua del Municipio de World Archaeology, pp. 538–554.
Palestina (Caldas). Proyecto de Rescate y Monitoreo Arqueológico del Aeropuerto del Piperno, D.R., 2006. The origins of plant cultivation and domestication in the Neotropics:
Café. Centro de Museos. Universidad de Caldas, Asociación Aeropuerto del Café, a behavioral ecological perspective. In: Kenett, D.J., Winterhalder, B. (Eds.),
Manizales, Colombia. Behavioral Ecology and the Transition to Agriculture. University of California Press,
Hurt, W., van der Hammen, T., Correal, C., 1977. The El Abra Rockshelters, Sabana de Berkeley, pp. 137–166.
Bogotá, Colombia, South America. Occasional Papers and Monographs No. 2. Indiana Piperno, D.R., 2011a. The origins of plant cultivation and domestication in the New
University Museum, Bloomington. World Tropics: patterns, process and new developments. Curr. Anthropol. 52 (4),
INTEGRAL, 1997. Arqueología de rescate: vía alterna de la troncal de Occidente río 453–470.
Campoalegre-Estadio Santa Rosa de Cabal. Medellín. INTEGRAL S.A., Ministerio de Piperno, D.R., 2011b. Prehistoric human occupation and impacts on Neotropical forest
Transporte, Instituto Nacional de Vías.Medellín (unpublished manuscript). landscapes during the Late Pleistocene and Early/Middle Holocene. In: Bush, M.B.,
Iriarte, J., 2007. New perspectives on plant domestication and the development of agri- Flenley, J.R., Gosling, W.D. (Eds.), Tropical Rain Forest Responses to Climatic
culture in the New World. In: Denham, T., Iriarte, I., Vrydaghs, L. (Eds.), Rethinking Change, second ed. Praxis, Chichester, pp. 185–206.
Agriculture: Archaeological and Ethnoarchaeological Perspectives. One World Piperno, D.R., 2017. Assesing elements of an extended evolutionary synthesis for plant
Archaeology, Left Coast Press, pp. 167–188. domestication and agricultural origin research. PNAS 114 (25), 6429–6437.
Iriarte, J., 2009. Narrowing the gap: exploring the diversity of early food-production Piperno, D.R., Pearsall, D.M., 1998. The Origins of Agriculture in the Lowland Neotropics.
economies in the Americas. Curr. Anthropol. 50 (5), 677–680. Academic Press, San Diego.
Isendahl, C., 2011. The domestication and early spread of manioc (Manihot esculenta Piperno, D.R., Dillehay, T., 2008. Starch grains on human teeth reveal early broad crop
Crantz): a brief synthesis. Latin Am. Antiquity 22 (4), 452–468. diet in northern Peru. PNAS 105 (50), 19622–19627.

171
F.J. Aceituno, N. Loaiza Journal of Anthropological Archaeology 49 (2018) 161–172

Piperno, D.R., Andres, T.C., Stothert, K.E., 2000. Phytoliths in Cucurbita and other pone.0062707.
Neotropical Cucurbitaceae and their occurrence in early archaeological sites from the Salgado, H., 1988–1990. Asentamientos precerámicos en el alto y medio río Calima,
lowland American tropics. J. Archaeol. Sci. 7 (3), 193–208. Cordillera Occidental de Colombia. Cespedesia 16–17 (57–58), 139–162.
Piperno, D.R., Ranere, A.J., Dickau, R., Aceituno, F.J., 2017. Niche construction and Salgado, H., 1995. El precerámico en el cañón del río Calima, cordillera occidental. In:
optimal foraging theory in Neotropical agricultural origins: a re-evaluation in con- Cavelier, I., Mora, S. (Eds.), Ámbito y Ocupaciones Tempranas de la América
sideration of the empirical evidence. J. Archaeol. Sci. 78, 214–220. Tropical. Fundación Erigaie. Instituto Colombiano de Antropología, Bogotá, pp.
Piperno, D.R., Moreno, J.E., Iriarte, J., Holst, I., Lachnlet, M., Jones, J.G., Ranere, A.J., 91–98.
Castanzo, R., 2007. Late Pleistocene and Holocene environmental history of the Santos, G., 2010. Diez Mil Años de Ocupaciones Humanas en Envigado Antioquia. El Sitio
Iguala valley, Central Balsas Watershed of Mexico. PNAS 104 (29), 11874–11881. La Morena. Alcaldía de Envigado, Envigado.
Piperno, D.R., Ranere, A.J., Holst, I., Iriarte, J., Dickau, R., 2009. Starch grain and phy- Santos, G., Monsalve, C.A., Correa, L.V., 2015. Alteration of tropical forest vegetation
tolith evidence for early ninth millennium B.P. maize from the central Balsas river from the Pleistocene-Holocene transition and plant cultivation from the end of early
valley, Mexico. PNAS 106 (13), 5019–5024. Holocene through middle Holocene in northwest of Colombia. Quat. Int. 363, 28–42.
Politis, G., 1995. Mundo Nukak. Fondo de Promoción de Cultura Banco Popular, Bogotá. Smith, B.D., 1998. The Emergence of Agriculture. Freeman, New York.
Politis, G., 1996. Moving to produce: Nukak mobility and settlement patterns in Smith, B.D., 2001. Low-level food production. J. Archaeol. Res. 9, 1–43.
Amazonia. World Archaeol. 27 (3), 492–511. Smith, B.D., Winterhalder, B. (Eds.), 1992. Evolutionary Ecology and Human Behavior.
Politis, G., 2007. Nukak Ethnoarchaeology of an Endagered Amazonian People. Left Coast Aldine de Gruyter, Hawthorne, New York.
Press, California. Sthal, P.W., 2016. Holocene biodiversity: an archaeological perspective from the
Posey, D.A., 1984. A preliminary report on diversified management of tropical forests by Americas. Annu. Rev. Anthropol. 25, 105–126.
the Kayapó Indians of the Brazilian Amazon. Adv. Econ. Bot. 1, 112–126. Stiner, M.C., Kuhn, S.L., 2016. Are we missing the “sweet pot” between optimality theory
Pringle, H., 1998. The slow birth of agriculture. Science 282, 1446–1450. and niche construction theory in archaeology? J. Anthropol. Archaceol. 44, 177–184.
Richerson, P.J., Boyd, R., Bettinger, R.L., 2001. Was agriculture impossible during the Tabares, D., 2004. Fase I: Prospección río Campoalegre, mundo arcaico en la región del
Pleistocene but mandatory during the Holocene? A climate change hypothesis. Am. Cauca medio, Colombia (unpublished results).
Antiq. 66 (3), 387–411. Trancho, G.J., Robledo, B., López-Buies, I., Fabián, F.J., 1996. Reconstrucción del patrón
Rindos, D., 1984. The Origins of Agriculture. An Evolutionary Perspective. Academic alimenticio de dos poblaciones prehistóricas de la meseta norte. Complutum 7,
Press, San Diego. 73–90.
Rival, L., 1998. Domestication as a Historical and Symbolic Process: wild gardens and Vavilov, N.I., 1992. Origin and Geography of Cultivated Plants. Cambridge University
cultivated forests in the Ecuadorian Amazon. In: Balée, W. (Ed.), Advances in Press, Cambridge.
Historical Ecology. Columbia University Press, New York, pp. 232–250. Vivo, M., Carmignotto, A.P., 2004. Holocene vegetation change and the mammal faunas
Rival, L., 2006. Amazonian historical ecologies. J. R. Anthropol. Inst. (N.S) 79–94. of South America and Africa. J. Biogeogr. 31 (6), 943–957.
Rodríguez, C., 1997. Rescate arqueológico sitios Los Arrayanes Pk 91+150 Villamaría- Watson, P.J., 2001. Food production, origins of. In: Bates, P. (Ed.), International
Caldas y El Pomo Pk 7+200 ramal a Manzanares, Fresno-Tolima. Informe Final. Encyclopedia of the Social & Behavioral Sciences. Elsevier Science Ltd, pp.
Empresa Colombiana de Petróleos (ECOPETROL), Bogotá (unpublished results). 5722–5728.
Rodríguez, C., 2002. El valle del Cauca Prehispánico. Procesos socioculturales antiguos en Winterhalder, B., Kennett, D.J., 2006. Behavioral ecology and the transition from hunting
las regiones neohistóricas del alto y medio Cauca y la costa Pacífica Colombo-ecua- and gathering to agriculture. In: Kenett, D.J., Winterhalder, B. (Eds.), Behavioral
toriana. Departamento de Historia Facultad de Humanidades Universidad del Valle, Ecology and the Transition to Agriculture. University of California Press, Berkeley,
Fundación Taraxacum, Cali, Colombia. pp. 1–21.
Rojas, S., Tabares, D., 2000. Aportes para una historia en construcción: arqueología de Zeder, M.A., Bradley, D.G., Emshwiller, E., Smith, B.D., 2006. Documenting domestica-
rescate en la doble calzada Manizales-Pereira-Armenia. INVIAS-CISAN, Bogota (un- tion: bringing together plants, animals, archaeology, and genetics. In: Zeder, M.A.,
published results). Bradley, D.G., Emshwiller, E., Smith, B.D. (Eds.), Documenting Domestication: New
Roullier, C., Duputié, A., Wennekes, P., Benoit, L., Fernández, V.M., Rossel, G., Tay, D., Genetic and Archaeological Paradigms. University of California Press, Berkeley, pp.
McKey, D., Lebo, V., 2013. Disentangling the origins of cultivated sweet potato 1–12.
(Ipomoea batatas (L.) Lam.). PLoS ONE 8 (5). http://dx.doi.org/10.1371/journal.

172

You might also like