1 - Piperno Neotropicos

19
The Origins of Plant Cultivation and Domestication in the New World Tropics: Patterns, Process, and New Developments Author(s): Dolores R. Piperno Reviewed work(s): Source: Current Anthropology, Vol. 52, No. S4, The Origins of Agriculture: New Data, New Ideas (October 2011), pp. S453-S470 Published by: The University of Chicago Press on behalf of Wenner-Gren Foundation for Anthropological Research Stable URL: http://www.jstor.org/stable/10.1086/659998 . Accessed: 28/01/2012 16:33 Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at . http://www.jstor.org/page/info/about/policies/terms.jsp JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms of scholarship. For more information about JSTOR, please contact [email protected]. The University of Chicago Press and Wenner-Gren Foundation for Anthropological Research are collaborating with JSTOR to digitize, preserve and extend access to Current Anthropology. http://www.jstor.org

Transcript of 1 - Piperno Neotropicos

Page 1: 1 - Piperno Neotropicos

The Origins of Plant Cultivation and Domestication in the New World Tropics: Patterns,Process, and New DevelopmentsAuthor(s): Dolores R. PipernoReviewed work(s):Source: Current Anthropology, Vol. 52, No. S4, The Origins of Agriculture: New Data, NewIdeas (October 2011), pp. S453-S470Published by: The University of Chicago Press on behalf of Wenner-Gren Foundation for AnthropologicalResearchStable URL: http://www.jstor.org/stable/10.1086/659998 .Accessed: 28/01/2012 16:33

Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at .http://www.jstor.org/page/info/about/policies/terms.jsp

JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range ofcontent in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new formsof scholarship. For more information about JSTOR, please contact [email protected].

The University of Chicago Press and Wenner-Gren Foundation for Anthropological Research are collaboratingwith JSTOR to digitize, preserve and extend access to Current Anthropology.

http://www.jstor.org

Page 2: 1 - Piperno Neotropicos

Current Anthropology Volume 52, Supplement 4, October 2011 S453

� 2011 by The Wenner-Gren Foundation for Anthropological Research. All rights reserved. 0011-3204/2011/52S4-0019$10.00. DOI: 10.1086/659998

The Origins of Plant Cultivation andDomestication in the New World Tropics

Patterns, Process, and New Developments

by Dolores R. Piperno

CA� Online-Only Material: Supplement A

The New World tropical forest is now considered to be an early and independent cradle of agriculture.As in other areas of the world, our understanding of this issue has been significantly advanced bya steady stream of archaeobotanical, paleoecological, and molecular/genetic data. Also importantly,a renewed focus on formulating testable theories and explanations for the transition from foragingto food production has led to applications from subdisciplines of ecology, economy, and evolutionnot previously applied to agricultural origins. Most recently, the integration of formerly separateddisciplines, such as developmental and evolutionary biology, is causing reconsiderations of how novelphenotypes, including domesticated species, originate and the influence of artificial selection on thedomestication process. It is becoming clear that the more interesting question may be the originsof plant cultivation rather than the origins of agriculture. This paper reviews this body of evidenceand assesses current views about how and why domestication and plant food production arose.

Introduction

It has long been recognized that numerous New World plantdomesticates—more than half, in fact, of all American cropsand many of the staples that supported indigenous peopleswhen Europeans arrived—originated in Neotropical forests(e.g., Harris 1972; Sauer 1950). In the past few decades, alarge corpus of archaeobotanical, paleoecological, and mo-lecular/genetic information has become available from Cen-tral and South America that has led to a significantly increasedunderstanding of the geography and chronology of tropicalfood production. This information has established the low-land Neotropical forest as an early and independent cradle ofagricultural origins. In a volume published in 1998, DeborahPearsall and I reviewed and synthesized the known archae-ological, paleocological, ethnographic, and molecular/genetic

Dolores R. Piperno is Senior Scientist and Curator of Archaeobotanyand South American Archaeology in the Program for Human Ecologyand Archaeology in the Department of Anthropology at theSmithsonian National Museum of Natural History (10th Street andConstitution Avenue, Washington, DC 20560, U.S.A. [[email protected]]). This paper was submitted 13 XI 09, accepted 9 XII 10, andelectronically published 4 VIII 11.

evidence bearing on the prehistory of Neotropical agriculture(Piperno and Pearsall 1998).

Since then, much more evidence has been generated fromall the contributing disciplines. New archaeological sites havebeen discovered, excavated, and analyzed using the full com-plement of now-available archaeobotanical techniques. Im-portant early sites that were first identified and studied morethan 30 years ago and for which little to no botanical infor-mation was available have seen reexcavation and applicationsof microfossil research. Phytoliths and in some cases mac-rofossils from other early human occupations have been re-visited during the past 10 years with the use of expanded andimproved modern reference collections, and some of theseremains have been directly dated. Starch grain data from stonetools and human teeth are becoming available from a growingnumber of sites, and construction of large modern referencecollections has allowed identification of some domesticatedroot and seed crops. Major collecting efforts of importanteconomic taxa such as the Cucurbitaceae have been under-taken, adding significant knowledge about modern landracediversity in Cucurbita and Lagenaria, along with distributionalranges of wild species. Last but not least, newer molecularand other information has refined and in some cases revisedthe geography of plant domestication and also caused us toreconsider how we should view and come to an understanding

Page 3: 1 - Piperno Neotropicos

S454 Current Anthropology Volume 52, Supplement 4, October 2011

of the domestication process. In this paper, I review the pres-ently available information and assess current views of howand why plant food production and domestication arose.1

The Geography of Domestication

Agriculture may originate in discrete centers or evolve overvast areas without definable centers.(Jack R. Harlan 1971:468)

It is useful to begin by assessing what is currently knownabout the geography of Neotropical plant domestication (fora summary, see fig. 1). Scholars such as David Harris, CarlSauer, Jack Harlan, and others of their generations knew thattropical forest agriculture probably developed independentlylong before European arrival, and they for the most partunderstood whether Central America (largely Mesoamerica)or South America was the original home of the major crops.More precise area origins for New World crops, and whethersingle or multiple domestications or even hybrid origins oc-curred in some, was often less clear for a number of reasons.Wild congenerics that potentially could be progenitor specieson the basis of shared morphological attributes or apparentlack of hybridization barriers were often broadly distributed,and they sometimes occurred in both Central America andSouth America (e.g., Manihot [manioc]; Cucurbita [squashesand gourds]; Zea [teosinte], the ancestor of maize; Ipomoea[sweet potato]; Xanthossoma [malanga or cocoyam, NewWorld representatives of the taro family]; Dioscorea [yams];Phaseolus lunatus [sieva beans]; Gossypium [cotton]; manytree crops; Piperno and Pearsall 1998).

The analysis of protein and DNA-based molecular markershas significantly elucidated these issues, most recently forsome major root, seed, and tree crops such as manioc, variousspecies of Cucurbita, chayote (Sechium edule), P. lunatus (sievabeans), Leucaena spp. (guaje), Spondias purpurea (jocote orMexican plum), Bactris gasipaes (pejibaye or peach palm),peanuts (Arachis hypogaea), South American cotton, and oth-ers (e.g., Bertoti de Cunha et al. 2008; Cross, Saade, andMotley 2006; Hughes et al. 2007; Leotard et al., forthcoming;Milla, Isleib, and Stalker 2005; Miller and Schaal 2005; Motta-Aldana et al., forthcoming; Olsen 2002; Olsen and Schaal1999; Robledo, Lavia, and Seijo 2009; Rodrigues, Filho, andClement 2004; Sanjur et al. 2002; Westengen, Huaman, andHeun 2005; fig. 1). Furthermore, it has been conclusivelyshown that the wild ancestor of maize is not native to thesemiarid Mexican highlands but rather to lower, warmer, andmoister habitats of Guerrero and Michoacan states, where

1. The following definitions are used in this paper. “Cultivation” and“farming” refer to the preparation of plots specified for plant propagationand repeated planting and harvesting in such plots. A “cultivated plant”or “cultivar” refers to those that are planted and harvested, regardless oftheir domesticated status. “Domesticated species” are those that havebeen genetically altered through artificial selection such that phenotypiccharacteristics distinguish them from wild progenitors.

seasonal tropical forest is the native vegetation (e.g., Doebley2004; Matsuoka et al. 2002; van Heerwaarden et al. 2011).

In many cases studied where multiple domestications werediscussed, single domestications are now indicated (e.g., man-ioc, maize, the important pejibaye palm; Leotard et al., forth-coming; Matsuoka et al. 2002; Olsen 2002; Olsen and Schaal1999; Rodrigues, Filho, and Clement 2004). On the otherhand, a double domestication of the lima bean, one event inthe southern Ecuador/northern Peru Andes leading to thelarge-size lima and another in the humid tropical lowlandsleading to the small-seeded sieva bean, is again indicated bythe newest molecular data (Motta-Aldana et al., forthcoming).Furthermore, lowland western Mexico is definitively revealedas an origin area for the sieva, and it is possible that anotherindependent domestication event occurred in the lowlands ofCentral or South America (Motta-Aldana et al., forthcoming).More than one domestication of an important Central Amer-ican tree crop, S. purpurea, also appears likely (Miller andSchaal 2005).

An Old World plant, the bottle gourd, was transported toand well used throughout tropical America during the pre-Columbian era. Erickson et al. (2005), working with modernand ancient DNA, established conclusively that prehistoricplants originated in Asia, not Africa, as was previously as-sumed. The archaeological rinds they studied bore the markof domesticated plants. The investigators concluded that hu-mans, not ocean currents, probably carried the domesticatedplant from Asia during the initial colonization of the NewWorld.

Importantly, molecular data have also elucidated in moredetail some of the different processes involved in the do-mestication of plants (see also “The How of Plant Domes-tication”). For example, interesting cases of hybrid origins ofMesoamerican polyploid tree crops were confirmed in theimportant genus Leucaena, whereby human-mediated sym-patry in “backyard gardens” of previously separated wild taxa,extensive predomestication cultivation, and subsequent spon-taneous hybridization led to the emergence of different do-mesticated species in the genus (Hughes et al. 2007). Themost widely cultivated Mesoamerican cactus, Opuntia ficus-indica, probably arose by the same means (Griffith 2004;Hughes et al. 2007). The importance of these processes andbackyard, also called “dooryard,” cultivation in contributingto the stock of plants domesticated prehistorically was pro-posed long ago by Edgar Anderson (1954). We should re-member that other major American crops, such as peanuts,are products of the hybridization of two different species andthe how, when, and where of the predomestication hybridi-zation events that led to their initial emergence are not wellunderstood.

Figure 1 shows known or postulated geographic zones ofdomestication for some Neotropical crops on the basis ofcurrent molecular, archaeological, and ecological evidence (fora more complete list of crops native to the Neotropics, seeHarlan 1992). In some cases, designations of the circled and

Page 4: 1 - Piperno Neotropicos

Piperno Plant Cultivation and Domestication in the New World Tropics S455

other localities on the maps as origin areas for particular cropsshould be treated as hypotheses that require testing with ad-ditional molecular and archaeological data. For example,more precise domestication localities for yam Dioscorea trifida,leren Calathea allouia, cocoyam Xanthossoma sagittifolium,and achira Canna edulis within their likely origin areas ofnorthern South America may be forthcoming once relevantmolecular studies have been undertaken. There is a majorroot crop for which neither Central American nor SouthAmerican origins have been conclusively demonstrated. Sweetpotato’s presumed wild ancestor is Ipomoea trifida, a poorlydemarcated taxon naturally distributed from Mexico throughnorthern South America (Piperno and Pearsall 1998). Somemolecular work based on amounts of diversity in modernlandraces suggests that it may be Central American in origin(Zhang et al. 2000), although the oldest archaeological evi-dence for it by far derives from western South America (Pip-erno and Pearsall 1998; Shady, Haas, and Creamer 2001).Centers of present diversity do not always accurately pinpointa center of origin, and additional molecular work focusingon the construction of a rigorous phylogeny of wild and do-mesticated sweet potato is needed to resolve the issue.

Even with the existing uncertainties, it is clear that croporigins were spatially diffuse. Using the conventional defi-nition of a center—a circumscribed region where agriculturebegan and out of where it spread—Mesoamerica, more pre-cisely Mexico, may still qualify, although it is clear now thatmaize, the common bean (Kwak, Kami, and Gepts 2009) andother Phaseolus species (the tepary bean), various tree crops,chile peppers, cotton, chayote, and others were domesticatedin different and sometimes ecologically dissimilar regions ofthe country (see also Piperno, forthcoming). In South Amer-ica, an attempt to designate a single circumscribed center orcore area of agriculture would clearly not be supported. Theorigins of major and now minor crops are spread from thenorthern parts to the southern parts of the continent, westand east of the Andes, mostly in seasonal types of lowlandtropical forest for major root and seed crops but also in low-land wet forests and midelevation moist forest habitats.

Patterns indicating multiple areas of domestication becomeeven more accentuated when highland crops such as Phas-eolus, Amaranthus spp., Chenopodium spp., and the variousAndean tubers are added to the picture (the Andean complexalso includes additional species of Cucurbita and Capsicum;Piperno, forthcoming; Piperno and Dillehay, forthcoming).Crops that would eventually become major staple foods orcondiments and that were commonly grown together afterfood production was established—such as maize, manioc, var-ious squashes, chile peppers, sweet potato, and Phaseolus andCanavalia beans—had spatially disparate areas of origin anddid not at first spread together. In fact, some of the earliestand most widespread anchors of lowland food production arenow minor, even disappearing, foods (below). Early patternsof dispersals probably did not involve significant populationmovements or diffusion of crops in packages.

Of course, knowing that crop plant origins were geograph-ically widespread does not necessarily lead to a conclusionthat food production was independently developed wherevera particular crop originated. Rather, the spread of a crop orcrops into new regions may have inspired receiving culturesto grow their own native plants. This issue is further com-plicated by the fact that we do not know when some of thecrops displayed in figure 1 were initially brought under cul-tivation. How many truly independent developments of low-land food production were there? Given that lowland north-ern and southern South American domestication zones areseparated by large distances and that a number of plants nativeto these areas were taken under cultivation and domesticatedby the middle of the eighth millennium BP, it is difficult tosee how the northern and southern lowland regions do notform at least two to three independent areas of food pro-duction (e.g., D1, D3, and D4 in fig. 1B). These questionswill be further illuminated in the near future as more dataare accumulated.

In sum, Harlan’s (1971) idea that peoples over a wide geo-graphic area were simultaneously engaging in early cultivationand domesticatory relationships with plants and considerablyinfluenced the early development of some domesticates afterthe plants left their native areas seems particularly relevant inthe light of current data. Opportunities for such kinds ofprocesses to occur would have been even greater if, as in theOld World (e.g., Fuller 2007; Jones and Brown 2007; Weiss,Kislev, and Hartmann 2006; Willcox, Fortnite, and Herveux2008), protracted periods of predomestication cultivationand/or spread of predomesticated crops occurred in the Neo-tropics. This issue is discussed in more detail below.

Finally, molecular and botanical studies together with anincreasing amount of archaeobotanical data, summarized be-low, tell us that the wild ancestors of many important cropplants are native to the seasonal tropical forest, those for-mations that receive a 4–7-month-long period every year dur-ing which little to no rain falls. Annual precipitation in theseareas averages about 1,200–1,800 mm a year, soils are lessweathered and thus more highly fertile than in ever-wet (asea-sonal) forest, and the prolonged dry season enabled earlyfarmers to efficiently clear vegetation and prepare plots forplanting with the simple use of fire. Seasonally dry tropicalforests do not carry the distinction of their rain forest relatives,but their prominent position in Neotropical agricultural or-igins is clear.

Early Food Production and Its Culturaland Ecological Contexts

The New World was colonized before 15,000 BP, by whichtime human populations had traversed most of the westernSouth American landmass, probably by moving along thePacific Coast, and found themselves in what is now southernChile (Dillehay et al. 2008). (All dates given in the text arein calibrated 14C years; uncalibrated and calibrated determi-

Page 5: 1 - Piperno Neotropicos

S456 Current Anthropology Volume 52, Supplement 4, October 2011

Figure 1. Postulated domestication areas for various tropical crops inCentral and South America. Open circles are archaeological and paleo-ecological sites with early domesticated crop remains. The numbers inparentheses after a taxon indicate that more than one independent do-mestication occurred. The possible area of origin for the sieva bean ex-tends into the Pacific lowlands north of the oval area. The oval here andthose in B labeled D1–D4 designate areas where it appears that morethan one or two important crops may have originated. Arrows point toapproximate areas and are not meant to denote specific domesticationlocales (for more details and sources used in the figure, see Piperno andPearsall 1998; Piperno 2006a; see also Cross, Saade, and Motley 2006;Hughes et al. 2007; Matsuoka et al. 2002; Miller and Schaal 2005; Motta-Aldana et al. forthcoming; Olsen and Schaal 1999; Plowman 1984; Ro-drigues, Filho, and Clement 2004; Sanjur et al. 2002; Wessel-Beaver 2000;Westengen, Huaman, and Heun 2005). Modern vegetation zone guidesare (a) 1, tropical evergreen forest; 2, tropical semievergreen forest; 3,tropical deciduous forest; 4, savanna; 5, low scrub/grass/desert; 6, mostlycactus scrub and desert; and (b) 1, tropical evergreen forest (TEF); 2,tropical semievergreen forest (TSEF); 3, tropical deciduous forest (TDF);4, mixtures of TEF, TSEF, and TDF; 5, mainly semievergreen forest anddrier types of evergreen forest; 6, savanna; 7, thorn scrub; 8, caatinga; 9,cerrado; 10, desert.

nations for specific dates from sites discussed are found intable 1.) The earliest indisputable evidence of human occu-pation in what is now the lowland Neotropical forest comesfrom sites distributed from Belize to eastern Brazil that werefirst occupied at about 13,000 BP (Piperno and Pearsall 1998;Ranere and Cooke 2003). Beginning soon after the Pleistocene

ended at about 11,400 BP, human settlement of the Neotropicsbegan to change from these sparsely distributed and short-term occupations. People “settled into” their landscapes, stay-ing for longer and/or more frequently returning to specificlocations, and they frequently manipulated and altered theirenvironments by creating clearings in forests and/or burning

Page 6: 1 - Piperno Neotropicos

Piperno Plant Cultivation and Domestication in the New World Tropics S457

Figure 1. (Continued)

them. They developed tool kits indicating that for the firsttime the exploitation of plants was as important an economicstrategy as hunting had been (e.g., Gnecco and Aceituno 2006;Mora 2003; Ranere and Cooke 2003; Ranere et al. 2009).Archaeobotanical information indicates that food productionbegan in a number of localities in tropical Central and SouthAmerica during the early Holocene (between 11,000 and 7600BP), not long after the Neotropical climate and vegetationunderwent profound changes associated with the end of thePleistocene (discussed in more detail below).

The best evidence currently comes from the Central BalsasValley of southwestern Mexico (Piperno et al. 2009; Ranereet al. 2009), central Pacific and western Panama (Dickau 2010;Dickau, Ranere, and Cooke 2007; Piperno 2006c, 2009; Pip-erno et al. 2000), the sub-Andean and premontane zones(elevation between 1,000 and 1,600 m) of the Cauca and Porcevalleys in Colombia (e.g., Aceituno and Castillo 2005; Brayet al. 1987; Gnecco and Aceituno 2006), the Colombian Am-azon (Cavelier et al. 1995; Mora 2003; Piperno and Pearsall1998), southwestern Ecuador (Pearsall 2003; Pearsall, Chan-dler-Ezell, and Chandler-Ezell 2003; Pearsall, Chandler-Ezell,and Zeidler 2004; Piperno 2009; Piperno and Stothert 2003;Zarillo et al. 2008), and the Zana Valley of northern Peru

(Dillehay et al. 2007; Piperno and Dillehay 2008). The veg-etation of all of these areas was humid tropical forest withthe exception of the Vegas, Ecuador, sites located at an ecotonebetween forest and scrub vegetation. Table 1 contains detailedinformation on crop plant occurrence and chronology. As-sociated published references and other information aboutthe sites involved can be found in CA� online supplementA. Figure 2 displays site locations. The reader should refer tothe citations given above and in CA� supplement A in dis-cussions that follow below.

The relevant archaeobotanical data are in the main partfrom microfossils, namely, starch grains recovered from nu-merous securely dated stone tools and human teeth and phy-toliths retrieved from the same stone tools and/or closelyassociated sediments. In several cases, phytoliths and starchgrains have been directly dated. Macrobotanical informationis available from Colombia and northern Peru, and paleo-ecological efforts allied with archaeological work provide pol-len, phytolith, and charcoal data indicating crop presence and/or substantial human vegetational modification near sites. Ta-ble 1 also includes early macrobotanical data from the aridcoast of Peru, where crops grown were largely from elsewhere

Page 7: 1 - Piperno Neotropicos

Table 1. Crop plant occurrence and chronology

Site Age Crop plants

Mexico:Guerrero State:

Xihuatoxtla Shelter By 7920 � 40 BP (by 8960–8940, 8850–8840,and 8780–8630 cal BP)

Maize (Phy and SG-GS), Cucurbita (Phy)

Tabasco State:San Andres 6208 � 47 BP (7204–6904 cal BP) Maize (Phy, Po)

Panama:Central Pacific Panama:

Aguadulce Rock Shelter By ca. 8600 cal BP Cucurbita moschata, leren, bottle gourd(Phy), Arrowroot (Phy, SG-GS)

6910 � 60 BP (7740–7640 cal BP) Maize, manioc (SG-GS),7061 � 81 BP (7922–7754 cal BP) Maize (Phy)By ca. 5700 cal BP Dioscorea trifida (SG-GS)

Cueva de los Ladrones 6860 � 90 BP (7804–7631 cal BP) Maize (SG-GS, Phy, Po)Cerro Mangote 6810 � 110 (7779–7584 cal BP) Maize (SG-GS)

Western Panama:Chiriqui Rock Shelters 6560 � 120 BP (7554–7381 cal BP) Arrowroot, maize (SG-GS)

Ca. 5600 cal BP Manioc (SG-GS)Hornito 6270 � 270 BP (7779–7584 cal BP) Maize (SG-GS)

Colombia:Middle Porce Valley Between 6280 � 120 and 5880 � 80 BP (be-

tween 7321–7032 and 6799–6597 cal BP)Maize (SG, Phy-GS; Po)

Middle Cauca Valley:El Jazmin 7590 � 90 BP (8493–8313 cal BP) Dioscorea (SG-GS [D. trifida?])

Between ca. 7000 and 5000 BP (ca. 8000–6000cal BP)

Maize (Po)

Middle Cauca Valley, Calima Region:El Recreo 7980 � 120 and 7830 � 140 BP C. cf. moschata (Phy); Persea americana

(M, [Cul?]);(9001–8674 and 8903–8508 cal BP) Cucurbitaceae (M)

Hacienda Lusitania 15150 � 180 BP (16138–5721 cal BP) Maize (Po)Hacienda El Dorado 6680 � 230 BP (7771–7349 cal BP) Maize (Po)

Upper Cauca Valley:San Isidro 9530 � 100 BP (11,058–10,706 cal BP) Bottle gourd (M, Phy), Cucurbita (Phy

[Cul.?]), P. americana (M), Maranta cf.arundinacea (SG-GS [Cul.?])

Colombian Amazon:Middle Caqueta Region:

Pena Roja 8090 � 60 BP (9107–8884 cal BP) Cucurbita, leren, bottle gourd (Phy)Abeja 14695 � 40 BP (15539–5351 cal BP) Maize, manioc (Po)

Southwestern Ecuador:Las Vegas Sites:

OGSE-80 and OGSE-67 Between 10,130 � 40 and 9320 � 250 BP(11,750–10,220 cal BP)

Cucurbita ecuadorensis (Phy)

9320 � 250 BP (11,060–10,950, 10,780–10,220cal BP)

Leren, bottle Gourd (Phy)

7170 � 60 BP (8015–7945 cal BP) Maize (Phy)15820 � 180 BP (6850–6810 cal BP) Maize (Phy)

Valdivia Sites:Real Alto Ca. 4300 BP (ca. 5000 cal BP) Leren, achira, arrowroot, maize, manioc

(SG, Phy-GS; Phy), jack bean, cotton(M)

Loma Alta 4470 � 40 BP (5260–5000 cal BP) Arrowroot, maize, manioc, jack bean,Capsicum (SG-Cer)

Ca. 5500–4400 BP (ca. 6500–5200 cal BP) Maize, Cucurbita, Achira (Phy)Ecuadorian Amazon:

Ayauchi Ca. 5300 BP (ca. 6000 cal BP) Maize (Po, Phy-Lake sediments)Eastern Amazon:

Geral, Brazil 5760 � 90 BP (6662–6464 cal BP) Slash-and-burn cultivation (?)Ca. 3350 BP (ca. 3800 BP) Maize (Po, Phy-Lake sediments)

Page 8: 1 - Piperno Neotropicos

Piperno Plant Cultivation and Domestication in the New World Tropics S459

Table 1. (Continued)

Site Age Crop plants

Northern Peru:Zana Valley 9240 � 50 BP (10402–10253 cal BP) C. moschata (M)

7840 � 40 BP (8630–8580 cal BP) Arachis sp. (M)5490 � 60 BP (6301–6133 cal BP) Cotton (M)Ca. 7500 BP (ca. 8500 cal BP) Manioc (M)8210 � 180 BP (9403–8784 cal BP) C. moschata, Arachis, Phaseolus, Inga

feuillei (SG-HT)7120 � 50 BP (7950 cal BP) Coca (Erythroxylum novagranatense var

truxillense; M)Siches 9533 � 65 BP (11015–10885 cal BP; BGS 2426)

and 9222 � 60 BP (10243–10306 cal BP; BGS2475)

Cucurbita (Phy)

Southern Coastal Peru:Paloma Ca. 7800 BP (ca. 8800 cal BP) Bottle gourd (M)

5070 � 40 BP (5900–5740 cal BP) Cucurbita ficifolia (M)By 5300–4700 BP (6500–5700 cal BP) Phaseolus lunatus, Cucurbita spp., guava

(Psidium guajava; M)Chilca 1 5616 � 57 BP (6440–6310 cal BP) P. lunatus (M)

By 4400 BP (5400 cal BP) Cucurbita, Achira, Jicama (Pachyrhizusahipa), jack bean

Quebrada Jaguay 7660 � 50 BP (8445–8395 cal BP) Bottle gourd (M)Southeastern Uruguay:

Los Ajos 4190 � 40 BP (4800–4540 cal BP) Maize, Phaseolus (SG-GS); maize, Cucur-bita (Phy)

Note. Dates indicate the first appearance of crops in the different contexts from each site studied. GS p recovered from grinding stones; HT precovered from human teeth; Cer p recovered from food residues in ceramic pots; if none of these context designations is listed, the botanicalremains were recovered from sediments. M p macrobotanical; Phy p phytoliths; SG p starch grains; Po p pollen. Bold printed 14C dates indicatethat the determinations were made directly on the botanical material. Other 14C dates are on other materials (usually wood charcoal, sometimesshell and human bone) from closely associated contexts. Laboratory numbers are listed for previously unpublished radiocarbon dates. Cul? puncertainty as to whether remains are from wild or cultivated/domesticated plants.

and their first appearance provides a minimum date for theirdomestication.

The earliest crops were Calathea allouia (leren) and Ma-ranta arundinacea (arrowroot), both grown for their tubers;Cucurbita moschata, Cucurbita ecuadorensis, and possibly Cu-curbita argyrosperma squash; bottle gourd; maize; manioc;peanuts; avocado; and pacay (Inga feullei), another tree crop.The first three—leren, arrowroot, and C. moschata squash—along with the bottle gourd are persistently present in north-ern South America and Panama between 10,200 and 7600 BP,underlining their probable northern South American origins.In many cases, different kinds of archaeobotanical remainsprovide mutually supporting evidence for the same crop spe-cies from the same site and specific context (table 1). It isobvious now that the earliest crop complexes were neitherseed, tree, nor root crop based but rather mixtures of thesedifferent elements.

During the middle eighth millennium BP, the first signs ofmajor crop movements northward from their area(s) of originin southern South American can be seen. Macrofossil andstarch grain data show that peanuts moved into the ZanaValley of northern Peru by 8500 BP. Macrofossils of manioc(identification confirmed by starch grains isolated directlyfrom root remains) also occur there by about 8500 BP, andstarch grains from the plant are present in central Panama at

7600 BP. Pollen evidence from the Colombian Amazon in-dicates that manioc arrived there before 5800 BP. It is possiblethat peanuts and manioc moved north together from a com-mon area of origin (fig. 1B). Chile peppers were well dispersedin southern Central America and South America by 6000–5000 BP (Perry et al. 2007).

Recent studies in the Central Balsas River Valley of Mexico,maize’s postulated cradle of origin, document the presenceof maize phytoliths and starch grains at 8700 BP, the earliestdate recorded for the crop (Piperno et al. 2009; Ranere et al.2009). A large corpus of data indicates that it was dispersedinto lower Central America by 7600 BP and had moved intothe inter-Andean valleys of Colombia between 7000 and 6000BP. Given the number of Cauca Valley, Colombia, sites thatdemonstrate early maize, it is likely that the inter-Andeanvalleys were a major dispersal route for the crop after it en-tered South America (table 1; fig. 2). Furthermore, directlydated starch grains from food residues in early Valdivia ce-ramics once again affirm maize and other crop presence indomestic contexts in these Early Formative (6000–5000 BP)occupations (Zarillo et al. 2008; see table 1 for other Valdiviacrop plant occurrences).

An abundance of artifactual evidence also speaks to earlytraditions of dedicated plant exploitation and cultivation.Stone implements used for plant processing (“edge-ground

Page 9: 1 - Piperno Neotropicos

S460 Current Anthropology Volume 52, Supplement 4, October 2011

Figure 2. Location of sites with early food production in Central andSouth America discussed in the text. The sites are shown with open circles.Sites with an open triangle are important examples of pre-6000 BP cropdiffusion into southern coastal Peru (Chilca 1, Paloma, and QuebradaJaguay) and regions of South America where agriculture as a whole hadbeen poorly documented (e.g., Geral, located in the interior of the Am-azon, and possibly Los Ajos, Uruguay, by ca. 4000 BP). They are notdiscussed in the text but are listed in table 1.

cobbles,” quern stone bases, hand milling stones) are commonin sites with early food-producing activities, and they are oftenpresent from the earliest Holocene onward in sites that wereoccupied at that time. Residues studied from these tools havecommonly produced starch grains and phytoliths from a va-riety of cultigens (table 1). In the Porce and Middle Caucavalleys of Colombia, the Colombian Amazon, and the ZanaValley in Peru, finely made stone hoes firmly dated to from9500–7300, 8000, and 7000 BP, respectively, occur in siteswith evidence of early food production (table 1) and in othersfor which botanical data are not available (Aceituno and Cas-tillo 2005; Gnecco and Aceituno 2006; Herrera et al. 1992;Mora 2003; Salgado 1995). The Cauca region Colombian ex-

amples, which by 7300 BP are beautifully polished, are amongthe most typical implements documented in early preceramicoccupations there. Stone axes presumably used for creatingopenings in the forest are also found in the Cauca, PorceValley, and Colombian Amazon occupations at the same time.

In the way of other economic remains documented, a num-ber of wild plant taxa, including a variety of palms and othertree fruits (e.g., Spondias, Erythrina, Byrsonima), yams otherthan Dioscorea trifida, Calathea species other than C. allouia,and Zamia have been identified. It is possible that some ofthem, particularly a few of the palm genera, were cultivated,but this cannot be empirically demonstrated with either themacrobotanical data or the microbotanical data. Current in-

Page 10: 1 - Piperno Neotropicos

Piperno Plant Cultivation and Domestication in the New World Tropics S461

formation similarly cannot be employed to adequately quan-tify changes in the proportions of wild and cultivated plantfoods as early periods of cultivation ensued. Bone often doesnot survive in the humid tropics, but in the few sites withfaunal records (e.g., Las Vegas and the Zana Valley), bothlarge and small animals were hunted. The totality of evidenceindicates broad-spectrum subsistence orientations shortly be-fore and at the beginning of food production.

With evidence accumulating rapidly now, other questionssuch as how early cultivated and domesticated plants moved(through movements of people, objects, or cultural knowl-edge) will take on increasing importance. This issue cannotbe discussed in any detail here, but considering the presentevidence, simple down-the-line forms of exchange that didnot involve significant population shifts or movements ofmaterial culture may best account for early crop plant dif-fusion. It should be stressed that different scenarios may havebeen true for later cases of crop diffusion through the Neo-tropics not discussed here, when Formative-period societieswere established and population numbers were much higherthan they were during the early and early middle Holocene.

It is clear that unlike in the Near East and China (see papersthis issue), Neotropical food production did not originate andtake hold in the context of large or fairly large permanentand nucleated villages situated in major river valleys. Rather,intensive foot surveys and excavations in the Balsas region ofsouthwestern Mexico, central and western Panama, south-western Ecuador, the Cauca and Porce regions, Colombia,and northern Peru show that between 11,000 and 7000 BP,sites are typically rock shelters and/or limited clusters of smallopen-air occupations that were located beside secondary wa-tercourses and seasonal streams whose small stretches of al-luvium likely were used for planting gardens. Settlements weretypically less than 1 ha in size, and many may have beenoccupied seasonally. Settlement organization was similar tomodern tropical hamlets and hamlet clusters, where one toa few nuclear families composed the residential community.

Early expressions of permanent settlements and seminu-cleated communities are found in the Zana Valley, Peru, whereoccupations between 10,000 and 7600 BP are small circularhouses located 200–400 m apart with stone foundations andstone-lined storage pits. The deep and dense Las Vegas, Ec-uador, middens, situated near the Pacific coast, where a widevariety of marine and estuary resources (e.g., mollusks, fish,and crabs) as well as plants and terrestrial animals were ex-ploited, may also have been occupied on a permanent basis.

It was not until substantially later in time that settlementsin these and other regions were positioned to exploit the richbottomland of significant river courses. The earliest such evi-dence comes from the Valdivia culture of southwest Ecuadorand dates to 5500 BP (e.g., Pearsall, Chandler-Ezell, andChandler-Ezell 2003; Raymond 2008). In other regions, thisdevelopment took place at about 4000–3400 BP. Another re-lated and often-discussed facet of early farming is the viewthat it should originate in zones of plentiful wild food re-

sources (see papers this issue). For a number of reasons, thehigh biodiversity of tropical forests does not translate intohabitats with abundant and stable wild resources for huntersand gatherers. Useful calorie-rich plant species occur in lownumber and are widely dispersed in space, and animal proteinis similarly far from plentiful, with the effect that naturalresources would generally be expected to support small groupsof mobile foragers (Piperno and Pearsall 1998:52–78). Re-source abundance would have been better, at least seasonally,in localities of early food production where permanent lakesoccurred, such as in an area of the Central Balsas, Mexico,studied recently (Piperno et al. 2007, 2009; Ranere et al. 2009).However, most early food producers did not have access tolakes. Piperno and Pearsall (1998:323) concluded that in theNeotropics, early farming occurred “in the most optimalzones of the most optimal types of forests and, in this sense,they represent resource abundance,” adding that “nonetheless,as long as people are not starving or otherwise experiencingfrequent and severe shortfalls of food, food abundance perse may have little relevance.” There is no reason to alter thisview.

When Did Neotropical Farming Becomea Significant Subsistence Practice?

The evidence increasingly indicates that it would be a mistaketo assume that Neotropical food production during the entire11,000–7000-BP period was a casual undertaking practicedby people who are better called foragers than farmers (seealso Iriarte 2007, 2009 for discussions of this issue). Multi-faceted archaeobotanical data—now including starch grainand phytolith evidence from the calculus of human teeth—along with settlement pattern, landscape modification (insome cases resulting from slash-and-burn cultivation), andartifactual information indicate that by 8800–7600 BP in theZana Valley, Peru; central Panama; probably the Central Bal-sas, Mexico; Tabasco, Mexico; and possibly other regions, asignificant number of dietary calories and nutrients camefrom crop plants. For example, more than 70% of the starchgrains recovered from the teeth of nine different Zana Valleyindividuals dated from 8800–7700 BP were from four crops:Phaseolus, Cucurbita moschata, peanuts, and Inga feuillea, atree (Piperno and Dillehay 2008). In Panama, the ColombianAmazon, and southwestern Ecuador, root crops such as lerenand arrowroot that were significant and reliable sources ofcarbohydrates are ubiquitous components of early crop plantassemblages. Cucurbita, another ubiquitous early plant, pro-vided high-quality proteins and fats. Furthermore, the starch,phytolith, and macrobotanical evidence from a number ofregions makes it clear that squashes (including the fruit fleshin the Zana Valley) were routinely consumed, had undergoneartificial selection for different traits related to fruit edibility,and thus were not primarily used as little-modified nondietaryplants.

In the Central Balsas and San Andres regions of Mexico

Page 11: 1 - Piperno Neotropicos

S462 Current Anthropology Volume 52, Supplement 4, October 2011

and in central Panama, forest clearance from slash-and-burncultivation is documented by pollen, phytolith, and charcoalrecords from lake sediment cores beginning at 7600–7200 BP(Piperno et al. 2007; Pohl et al. 2007; Pope et al. 2001). Smallirrigation canals are present in the Zana Valley beginning at7700 BP (Dillehay, Eling, and Rossen 2005). In a number ofregions, then, agricultural intensification began before 7000BP. In central Panama; the Zana Valley, Peru; and the Caucaand Porce regions, Colombia, it is also evident that site num-ber and artifact density significantly increased around 7600–7000 BP. These trends are likely the result of an increase inhuman carrying capacity made possible by effective systematicfood production.

In summary, these people appear to have been committedhorticulturists and slash-and-burn cultivators who, while stillintegrating planting with collecting and hunting, were takingimportant steps along the pathway to full-scale agriculture.The appearance of large sedentary and nucleated villages,which postdates 6000 BP throughout the Americas, shouldno longer be considered a necessary backdrop for the occur-rence or recognition of effective and productive agriculturein the Americas. This is all the more true when it is consideredthat indigenous Neotropical farmers still live in—and morecommonly did so in the recent past—small seminucleatedand shifting communities. The idea that the appearance ofsedentary village life and all of its trappings should be themeasure of when farming began in the Americas emergedfrom faulty comparisons with Near Eastern pre- and post-agricultural trajectories, which were a product of ecologicaland demographic circumstances very different from those as-sociated with the beginnings of Neotropical food production(Piperno and Pearsall 1998). As also stressed by Iriarte (2007),Vrydaghs and Denham (2007), and Denham (2011), agri-cultural origins in different parts of the world should be stud-ied using the totality of evidence relevant to the region beinginvestigated.

The How of Plant Domestication

Phenotypic innovation depends on developmental inno-vation, and developmental innovation spans a broader fieldof possibilities than does mutation alone.(Mary Jane West-Eberhard 2003:144)

Allowing for cryptic variants and novel phenotypes fromnew epistatic combinations to arise during domestication,it is easy to imagine that maize was domesticated fromteosinte.(John Doebley 2004:56)

Evolutionary Development, Gene Expression, andDevelopmental Plasticity

Recent developments in evolutionary biology and archaeo-botanical data gathering and interpretation indicate the need

to reassess how plant domestication occurred. Fundamentalnew insights about the origin of novel traits and adaptiveevolution are arising from the fields of evolutionary devel-opmental biology (evo-devo), molecular biology, and devel-opmental plasticity. In addition, a constant stream of detailedarchaeobotanical investigations has altered our views aboutthe trajectory of plant domestication following the advent ofsystematic cultivation (defined here as the cycle of plantingand harvesting in plots prepared for this purpose). I start firstwith new insights from the nonarchaeological spheres of re-search.

Gene regulation/expression acting during an organism’s de-velopment and phenotypic (developmental) plasticity are nowwidely viewed as significant forces, indeed by some investi-gators as the primary forces, in evolutionary diversificationand the origin of novel traits (e.g., Carroll 2005; Kruglyakand Stern 2007; Pigliucci 2005; West Eberhard 2003). Thereis good reason to believe that these research foci should be-come standard elements of domestication studies (Gremillionand Piperno 2009a). Regulatory genes are not protein-codingor functional genes but rather act to switch other genes onor off or change when and where in an organism they areactive or increase/decrease their effects, usually during earlyontogeny. This is the process of gene expression (e.g., Krug-lyak and Stern 2007). Novel phenotypes may then often resultfrom reorganizations of existing genomes, not by the spreador appearance of mutations, and new phenotypic variationcan rapidly arise in populations without a corresponding ge-netic change.

Developmental plasticity, recently given a complete treat-ment in a major book (West-Eberhard 2003), refers to theinherent capacity of organisms to rapidly produce novel her-itable phenotypes through one of several available develop-mental pathways in direct response to changes in their en-vironments. This capacity should be particularly importantin plants, which cannot simply get up and move to placesmore to their liking when physical and biotic conditionschange and become less favorable. Gene expression duringplant development orchestrates this process, resulting in dif-ferent phenotypic pathways to adulthood. The new pheno-types have the potential to become fixed (stabilized) by geneticassimilation/accommodation if the new ecological conditionsare maintained over multiple generations (West-Eberhard2003). An important source of genetic variation that mayenable the creation of novelties through developmental-mediated mechanisms is “cryptic genetic variation,” so namedbecause it is unexpressed in the phenotype of standing pop-ulations and thus normally hidden from selection (e.g., Gib-son and Dworkin 2004; Lauter and Doebley 2002). However,it may be set in motion or “released” by interactions betweengenes (epistasis) or by perturbations from the external en-vironment, rapidly resulting in new phenotypes. Lauter andDoebley (2002) have shown that maize’s wild ancestor pos-sesses a significant degree of cryptic variation that is associatedwith important traits such as polystichous (many-rowed) cobs

Page 12: 1 - Piperno Neotropicos

Piperno Plant Cultivation and Domestication in the New World Tropics S463

and nonshattering ears. Studies of this kind are needed inmany other crop plants.

Crop evolution has usually been characterized using theassumption that phenotypic change is driven by selection forrare mutants that are deleterious in wild plants or by selectionfor new random mutations that appeared after cultivationbegan. Associated single trait–single gene models are alsodeeply rooted in domestication studies. However, it is in-creasingly recognized that (1) inheritance of domesticatedtraits is often complex (simple Mendelian segregation is notindicated) and does not involve straightforward human se-lection on mutations that occurred as rare variants in wildancestral or early cultivated populations; (2) some “domes-tication genes” controlling crucial phenotypic attributes arein fact regulatory genes, for example, in maize (tb1 and tga1,controlling plant architecture and the formation of nakedgrains, respectively), rice (sh4 and qSH1, controlling shatter-ing), and tomato (fw2.2, underwriting fruit size increase); (3)these and other identified genes control developmental path-ways in specific immature tissues and organs that lead to adultphenotypic variability; (4) phenotypic outcomes are signifi-cantly influenced by the external environment as well as byregulatory and other gene-gene interactions; and (5) consid-erable cryptic variation occurs in at least one crop progenitor,teosinte, that provides preexisting genetic material that cantranslate into multiple developmental trajectories (e.g., Doe-bley 2004, 2006; Jones and Brown 2007; Lauter and Doebley2002).

Doebley (2004) describes what happens in crosses betweenteosinte and maize as a result of gene actions that alter thetrajectories of plant development: “The dichotomies of singleversus paired spikelets, shattering versus nonshattering ears,soft versus hard glumes, and two- versus multiranked earsare striking when one compares maize and teosinte. However,in F2 families, these discrete classes blur into a continuum ofphenotypes, and novel phenotypes and interactions appear”(54). All of this means that the domestication process involvedsimple unconscious/conscious selection targeted at specificmutations controlling discrete (single) traits and more com-plex plant-people-genetic-environmental interactions that of-ten played out during plant ontogeny and involved humans“reconfiguring” (Lauter and Doebley 2002:341) the preexist-ing pool of variability into new combinations. Present evi-dence suggests that in some crops, the latter may have beenmore important than the former. Gene expression appears tohave been very important, and it probably gave early culti-vators a great deal of phenotypic variation to work with insome crops and traits, especially as hybridization betweencrops and wild ancestors was initially common. This may haveeither slowed down or speeded up the domestication process.Furthermore, preexisting variation for plasticity would haveenabled crops to more quickly adapt to new ecological cir-cumstances as they were brought from their natural habitatsinto cultivated fields and dispersed out of their areas of origin,and it likely in part determined the amount of landrace di-

versification that was possible after domestication. Maize is aclassic example of a domesticated species that was capable ofadapting successfully to many different environmental cir-cumstances, and prehistoric peoples across the New Worldcreated many different races from it, neither surprising inlight of how much variation and plasticity are present in itswild ancestor.

Additional research will provide more detailed informationabout how gene expression and multiple possible develop-mental trajectories influenced the creation of domesticatedphenotypes, and this type of work will be important for an-imal domestication as well (e.g., Dobney and Larson 2006).What has so far been largely neglected in domestication stud-ies and should be an important part of future investigationsis the role of what West-Eberhard (2003) calls “environmentalinduction” in creating new phenotypes from ancestral pop-ulations (Gremillion and Piperno 2009a). This is a crux ofthe phenotypic plasticity argument as developed by West-Eberhard—that alternative phenotypes can be induced di-rectly by environmental change and subsequently geneticallyassimilated/accommodated (genetically stabilized) under theproper conditions. It is known, for example, that maizelikephenotypes in plant architecture can be induced in teosinteby environmental stresses such as lowered temperatures andlight intensity (West-Eberhard 2003). Given that crop do-mestication first occurred shortly after the environmental per-turbations that marked the end of the Pleistocene, humanenvironmental modification was significant starting duringthe early Holocene, and human care for/selection of novelphenotypes in secure ecological niches—cultivated fields—would lead to the genetic assimilation necessary to stabilizethe new phenotypes, the role of natural- and human-inducedenvironmental induction as a substitute for and/or comple-ment to artificial selection in engineering some of the majorphenotypic steps in crop evolution should be investigated (seeGremillion and Piperno 2009a for possible examples relatingto Chenopodium seed attributes and maize branching andinflorescence architecture).

Predomestication and Nondomestication Cultivation

Newer and more highly detailed archeobotanical informationindicates that rather than a rapid appearance of domesticatedplants shortly after cultivation began, protracted periods ofpredomestication cultivation (PDC) and even instances ofnondomestication cultivation (NDC) of cereals and pulsesoccurred in the Old World. PDC goes hand in hand withwhat appears to have been a nonsimultaneous developmentof the suite of traits that make up the domestication syndrome(e.g., Cohen 2011; Dillehay et al. 2007; Fuller 2007; Pipernoand Dillehay 2008; Weiss, Kislev, and Hartmann 2006; Weissand Zohary 2011; Willcox, Fortnite, and Herveux 2008; Zeder2011). Some important implications from the accumulateddata are obvious. Domestication took longer after cultivationwas initiated and was a more complex process than was

Page 13: 1 - Piperno Neotropicos

S464 Current Anthropology Volume 52, Supplement 4, October 2011

thought, involving different kinds and thresholds of (oftencompeting) natural and artificial selection pressures from re-gion to region and cultivar to cultivar. The cultivation ofmajor crop species may have arisen independently more thanonce. Productive and stable food production could be basedon morphologically wild plants. And if, as seems likely, PDCplants spread, defining a single geographically localized areaas the cradle of origin for the domesticated plant would provedifficult because multiregional processes involving gene flowwere at work (Allaby, Fuller, and Brown 2008; Brown et al.2009; Jones and Brown 2007).

I previously suggested that the commonplace focus on do-mestication as the preeminent event in human/plant rela-tionships is perhaps misplaced and that the crucial shift wemay want to understand is the origins of plant cultivation,when people began to repeatedly sow and harvest plants inplots prepared for this purpose (see Piperno 2006a for a moredetailed discussion). Given the evidence brought to light morerecently, I see no reason to change this view. Less attentionhas been paid to PDC and NDC issues in the New World,but some highland and lowland data already suggest that theywere components of early food production. At Guila NaquitzCave, highland Mexico, a type of morphologically wild runnerbean (Phaseolus spp.) was present between ca. 10,600 and8500 BP in quantities that suggested to the investigators thatpeople were artificially increasing their density by cultivatingthem (Flannery 1986). The plant was never domesticated. Inthe Zana Valley, Peru, hulls of peanuts (the nuts themselvesdid not survive) dated to 8500 BP do not exhibit some featuresof modern domesticated plants, nor do they conform to aknown wild species (Dillehay et al. 2007). Predomesticationcultivation of this plant and its transport out of its area oforigin in southern South America before it acquired the fullpackage of domesticated characteristics is thus indicated. Pea-nut nut starch grains recovered from human teeth, on theother hand, are identical to those from modern varieties ofArachis hypogaea and are unlike starch in modern wild pea-nuts closely related to A. hypogaea analyzed so far (Pipernoand Dillehay 2008). More work is needed on wild peanutstarch, but the suite of phenotypic traits that characterizedomesticated peanuts may not have developed all at once.

In fact, as discussed in detail elsewhere (Piperno 2006a:162–163; Piperno and Pearsall 1998), PDC and NDC sce-narios for agriculture, even if they were not called that, havelong seemed to a number of investigators to be particularlywell suited to the Neotropics (e.g., Harlan 1992; Harris 1989).Even today, many horticultural plots still contain morpho-logically wild plants that do not change their phenotypic char-acteristics even after many years of persistent cultivation.Crops grown for their belowground organs (e.g., Calatheaand Dioscorea spp.) and many palms and other fruit trees areamong the most stubborn taxa. Instances of PDC and NDCshould be sought for more widely in the New World by usingmultifaceted archaeobotanical data sets and allied paleoeco-logical evidence whenever possible.

Coming full circle back to the discussion that initiated thissection, the complex inheritance of and multifactorial influ-ences on domestication traits may have played roles in slowingdomestication. Furthermore, the effects of gene expressionand possible environment-on-phenotype influences at the on-set of plant cultivation and after may have led to pools ofinteresting phenotypic diversity in early crops that peoplepicked through and discouraged or encouraged over a pro-tracted period before the domesticated examples that we iden-tify archaeologically emerged.

The Why of Food Production

Human Behavioral Ecology and Agricultural Origins

Piperno and Pearsall (1998) and Piperno (2006a) offered anaccount of why and when Neotropical agricultural originsoccurred that is rooted in human behavioral ecology (HBE),especially optimal foraging theory, and they suggested thatHBE models had potential relevance in other regions of theworld. Explanations such as population pressure and changingsocial relationships (e.g., competitive feasting) used by someinvestigators to account for agricultural beginnings elsewhere(Piperno and Pearsall 1998:10–18) did not and still do notappear to be important in the Neotropics. Before discussingthe relationships between agricultural origins, HBE, and otherexplanations for the transition to food production, I shouldreiterate my belief that the near synchroneity of food pro-duction origins in at least seven widely dispersed and eco-logically disparate regions of the world when or shortly afterthe world’s fauna and flora were experiencing profound shiftsdriven by the end-Pleistocene ecological perturbations shouldcause us to look for common underlying processes that mayhave been influential in the transition wherever it occurred.Put at its simplest, it is difficult to believe that near-synchro-nous origins of such a major economic transformation werea coincidence, an accidental convergence of disparate region-ally specific underlying factors in widely dispersed and dif-ferent social systems. This is not to say that identifying specificaspects of Neolithic developments on a region-by-region basisis not important, only that we should in a deliberate way lookfor underlying commonalities rather than immediately em-phasizing putatively unique, supposedly causal local or re-gional variables that may turn out to be more apparent thanreal.

The basic concepts and uses of HBE in archaeology are bynow well known and discussed and will not be reviewed herein any detail (for recent literature reviews, see Bird andO’Connell 2006; Lupo 2007). In brief, HBE is concerned withthe adaptive plasticity of the human phenotype in responseto variation in its particular ecological and social environ-ment. It assumes that behavior is shaped by evolutionaryforces and asks “why certain patterns of behavior haveemerged and continue to persist” (Bird and O’Connell 2006:

Page 14: 1 - Piperno Neotropicos

Piperno Plant Cultivation and Domestication in the New World Tropics S465

144). The emphasis on a flexible phenotype means that HBE’sengine for change resides in human decision making. HBE isembedded in archaeological theory and empirical testing toa degree such that it is viewed along with dual inheritancetheory as a primary and distinct research tradition in evo-lutionary archaeology (e.g., Shennen 2008). As such, it hasbeen and will continue to be a focus of discussions, bothincisive and off the mark. Examples of the latter (Smith 2009;Zeder and Smith 2009) make faulty claims that explanationsderived from HBE are analogous to covering law models orthat HBE ignores the active roles of foragers in altering en-vironments (see Gremillion and Piperno 2009b for additionalcomments). Investigators in fact are drawn to HBE by itsfocused, hypothesis-driven, problem-solving approaches that,as mentioned, incorporate human agency and dynamic hu-man/environmental relationships. These elements constitutegood science and are productive means for “eliminating prob-lematic answers and identifying and pursuing more promisingones” (Bird and O’Connell 2006:171).

Although HBE has been used to study a broad array ofeconomic and social issues, “foraging theory” applicationsthat address subsistence decisions are most common to thispoint. At the heart of foraging theory is the assumption that,all things being equal, more efficient food procurement strat-egies should be favored by natural selection over those lessefficient, a largely unquestioned premise in nonhuman animalstudies. The simplest version of foraging models is called the“optimal diet” model, or the “diet breadth” model (DBM).It employs a straightforward currency—energy—to measurethe costs and benefits of alternative resource sets and assumesthat humans will have a goal of optimizing the energeticreturns of their subsistence labor. Recent research indicatingthat spatial memory in women is preferentially engaged andmost accurately expressed for calorie-rich foods further but-tresses the likelihood that natural selection shaped a humanproclivity for efficient plant foraging (New et al. 2007). Asreviewed in detail elsewhere (Piperno 2006a; Piperno andPearsall 1998:15–18), there are a number of good reasons whythe DBM can productively be applied to agricultural originsand dispersals. For one thing, its sometimes counterintuitivepredictions regarding resource choice bring new insight toissues such as the onset of “broad-spectrum” subsistence strat-egies and the links between diet breadth, technology, andresource intensification. Another strength of the model is itsability to elucidate from patterns in the archaeological recordimportant processual questions relating to human economicdecision making and its determinants vis-a-vis the focal pointof energetic constraints. Importantly, in this regard, paleo-ecological data can serve as objective informants on past shiftsin the availability of different subsistence resources and asproxies of changing returns to labor from foraging, addingvaluable information that can be used in conjunction with orin the absence of archaeological subsistence records.

A range of ethnographic information indicates that in theNeotropical forest, as elsewhere, energetic efficiency is a major

influence on food procurement decisions (reviewed in Pip-erno and Pearsall 1998). The following lines of evidence wereused to argue that optimal foraging and the relative energeticefficiency of resource sets available to foragers during the latePleistocene and early Holocene played a major role in Neo-tropical agricultural origins (details in Piperno 2006a; Pipernoand Pearsall 1998). First, a large set of paleoecological dataranging now from the Central Balsas region of southwestMexico to Bolivia (e.g., Burbridge, Mayle, and Killeen 2004;Piperno et al. 2007) indicates that the shift from foraging tofood production began within contexts of rapid and signifi-cant changes of climate, vegetation, and fauna occurring atthe close of the Pleistocene. Large increases in temperature,precipitation, and atmospheric CO2 levels resulted in vege-tational transitions from savanna-like/thorny scrub growth toseasonal types of tropical forest across the Neotropics. Manynow-extinct megafauna were replaced by the smaller, fewer,and more dispersed tropical forest fauna found in modernenvironments.

Second, on the basis of robust sets of ethnographic andecological data, these environmental perturbations wouldhave significantly lowered the overall efficiency of food pro-curement for hunters and gatherers in those zones whereopen-land types of vegetation gave way to forest when thePleistocene ended. For example, the big game and open-landplants that disappeared were almost certainly higher-rankedresources compared with those offered by the tropical forest,in which animals were far fewer and smaller, carbohydrateswere limited and spatially dispersed, and many plants weretoxic and required extensive processing before they were con-sumed. Decreasing foraging efficiency as dietary breadth ex-panded to incorporate these low-ranked resources was prob-ably an important selection pressure acting on human foodprocurement strategies during the early Holocene.

Third, on the basis of these factors, along with the costsof plant cultivation estimated from modern small-scale trop-ical farming, it appears that early Neotropical food-producingstrategies probably were more energetically efficient, not morecostly, than full-time foraging (see Piperno 2006a for a de-tailed discussion of this issue for the tropical forest). Finally,following the DBM, people would have initiated the culti-vation of some plants when the net return from this strategyexceeded the return from full-time hunting and gathering. Inlight of shifts in vegetational formations and available re-sources indicated by paleoecological records, the ca. 11,000–9000-BP time period should have been highly relevant for theinitiation of food production. It should be noted that all ofthe above points are especially relevant to areas that todaysupport or would support in the absence of human distur-bance highly seasonal types of tropical forest, where the end-Pleistocene environmental perturbations would have im-pacted foraging return rates most strongly.

Comparing these points with archaeological data (seeabove; Piperno and Pearsall 1998), it appears that predictionsof optimal foraging and the DBM are well supported and

Page 15: 1 - Piperno Neotropicos

S466 Current Anthropology Volume 52, Supplement 4, October 2011

effectively account for when and where food productionemerged in the lowland Neotropics. As has been discussed indetail elsewhere, other issues such as why some plants andnot others out of the many available were singled out forhuman manipulation can be elucidated using HBE (Piperno2006a; see below). The point made above about early farmers’energetic efficiency should be stressed. Although many in-vestigators assume that early farmers everywhere experienceddiminishing returns to their labor, studies in the southwesternand eastern United States and the Near East have also failedto support that notion when it was properly tested usingappropriate measures of resource costs and benefits (e.g., Bar-low 2002; Gremillion 2004; Simms and Russell 1997). Theenergetic efficiency of early farming compared with the lasthunting and gathering should be reevaluated in other regionsof the world. Especially in the cases where foraging theoryappears to effectively account for the initiation of food pro-duction, it would not be surprising to see a revision of pre-viously assumed cost/benefit equations for foraging versusfarming.

The DBM or combinations of DBM and other HBE theorieshave recently been productively applied to predict under whatcircumstances and how intensively foragers should becomefarmers in a number of world regions (for other examples,see, e.g., Barlow 2002; Bird and O’Connell 2006; Gremillion2004; Kennett and Winterhalder 2006). In the eastern UnitedStates (Gremillion 2004), it was found that some of the pre-dictions of the DBM (e.g., that immediate resource returnsprovide the best measure of their utility) did not fit the ar-chaeological data. In this case, the research led to a bettergrasp of the variables most important in agricultural transi-tions (e.g., risk reduction, energetic considerations of pro-cessing certain key resources with regard to the seasonal avail-ability of others), and refinements in how HBE models shouldbe applied in this regional circumstance. Students of HBE seethis as one of its strongest features. If a good fit betweentheory and data is not evident, then one moves on to find abetter-supported theoretical approach.

HBE and Risk

HBE is a highly flexible research program by no means limitedto simple energetic efficiency models and currencies. For ex-ample, risk sensitivity, an often-used explanation for why for-agers turned to farming, is increasingly being incorporatedinto HBE research (see papers in Kennett and Winterhalder2006). It is evident that under a number of circumstances,energetically efficient diets are also risk-minimizing diets andthat attention to risk avoidance does not necessarily overrideattention to energetic efficiency. It is worth repeating herethat there is also considerable ethnographic evidence thatamong tropical foragers and horticulturists, food sharing inthe context of extensive household exchange is the most im-portant strategy for mitigating risk (see Piperno and Pearsall1998:239–241 for a detailed discussion of these issues).

HBE and Coevolution

There is little disagreement that coevolutionary forces andfood production origins were entwined. As the general processunfolded, dependency of humans on certain plants increasedand vice versa, and both the plants and people involved ex-perienced fitness increases. However, when coevolution is op-erationalized at more detailed levels of analysis, discordancebetween the model and empirical archaeobotanical data be-comes evident. This was initially discussed by Piperno(2006a), and more examples can now be added.

The Cucurbitaceae and Marantaceae, each with importantexamples of early domesticated plants, are large families withmany edible species that respond to a human presence byquickly colonizing disturbed areas near living places. Thismakes them particularly suitable for testing the most completeand prominent explication of how agricultural transitions ex-emplify coevolution, that of Rindos (1984). The fruit rindsof most Cucurbitaceae genera produce high numbers of rec-ognizable phytoliths, and Marantaceae seeds and tubers aresimilarly rich in diagnostic phytoliths. These durable micro-fossils provide a record of exploitation unbiased by the pres-ervation factor. Marantaceae tubers also produce many di-agnostic starch grains, which would be expected to occur onstone tools that were used to process them. The sites of Xi-huatoxtla, Mexico; San Isidro, El Recreo, and Pena Roja, Co-lombia; Siches, Ecuador; and Zana Valley, Peru, can be addedto those from central Panama and Vegas, Ecuador, as examplesof where only Cucurbita (often along with bottle gourd) isrecorded among plants from this family. In the sequencesfrom Tehuacan and Guila Naquitz, Mexico, where macro-botanical preservation was excellent, one additional genus ofthe Cucurbitaceae, Apodanthera, occurred with Cucurbita andbottle gourd, but rarely, and it was considered a possiblemodern intrusive. Students of foraging theory would im-mediately notice that the fruits and seeds of wild Cucurbitaare among the largest in the family and therefore would likelyprovide higher return rates relative to other cucurbit taxa.

A similar example of highly selective exploitation is foundwith regard to the Marantaceae. Maranta arundinacea and afew species of Calathea, usually solely Calathea allouia, arethe only Marantaceae taxa represented in the tropical sites.There are numerous plants from a variety of other familieswith edible seeds and underground organs that should bevisible in phytolith and starch grain records had they beenexploited early on with any persistency, but they are not re-corded. Moreover, just a few species of wild grasses out ofthe hundreds that were available to human foragers were ma-nipulated in pre-Columbian Mexico: two species of Setaria(Austin 2006) and Zea mays ssp. parviglumis (Balsas teosinte).Only the latter was demonstrably domesticated. A correlationbetween grass domestication and seed size is also evident, asteosinte has the largest grains of any annual Mexican grass.

Therefore, Neotropical archaeobotanical data better sup-port predictions from foraging theory that stress the impor-

Page 16: 1 - Piperno Neotropicos

Piperno Plant Cultivation and Domestication in the New World Tropics S467

tance of deliberate directed human choice rather than theslowly unwinding reciprocal plant/human interactions in-volving experimentation with numerous different taxa thatwould be characteristic of coevolution. Efficiently procuringfood, especially in the tropical forest, would necessarily entailpaying serious attention to and deriving the most plant cal-ories, proteins, and fats from a relatively small set of theavailable wild species. Long periods of experimentation andmutualistic relationships with a high diversity of species donot appear to have been associated with squash, Marantaceae,and grass domestication in part because energetic efficiencywas probably a major determinant of subsistence decisions.

Conclusions

Discussions regarding the origins and dispersals of agricultureare increasingly based on robust empirical information as newand more precise data from all the contributing disciplinesof study become available. With gene expression, phenotypicplasticity, and HBE among the important additions to thehow and why of food production, the level of analysis hasbeen raised to incorporate the phenotype of both the plantsand humans involved. Given the immense importance thatgene expression and developmental plasticity are finding inevolutionary biology, it is likely that they will contribute muchto our understanding of how people domesticated plants.Human behavioral ecology similarly is likely to have growingimportance in agricultural origin studies. Microfossil datahave been shown to be essential lines of evidence in docu-menting the transition from foraging to farming in the Neo-tropics. As new sites are discovered and older ones reexam-ined, much more information can be expected on earlyagriculture in the New World’s tropical regions.

Acknowledgments

I thank Doug Price and Ofer Bar-Yosef for inviting me toattend the Temozon conference. I learned much from thestimulating discussions and papers presented there.

References CitedAceituno, F. J., and N. Castillo. 2005. Mobility strategies in Co-

lombia’s middle mountain range between the early and middleHolocene. Before Farming 2005(2), art. 2.

Allaby, R., D. Q. Fuller, and T. A. Brown. 2008. The genetic expec-tations of a protracted model for the origins of domesticated crops.Proceedings of the National Academy of Sciences of the USA 105:13982–13986.

Anderson, E. 1954. Plants, man, and life. London: Melrose.Austin, D. F. 2006. Fox-tail millets (Setaria: Poaceae): abandoned

food in two hemispheres. Economic Botany 60:143–158.Barlow, R. K. 2002. Predicting maize agriculture among the Fremont:

an economic comparison of farming and foraging in the AmericanSouthwest. American Antiquity 67:65–88.

Bertoti de Cunha, F., P. M. Nobile, A. A. Hoshino, A. Moretzsohn,M. de Carvalho, C. A. Lopes, R. Catalina, and M. A. Gimenes.

2008. Genetic relationships among Arachis hypogaea L. (AABB)and diploid Arachis species with AA and BB genomes. GeneticResources and Crop Evolution 55:15–20.

Bird, D. W., and J. F. O’Connell. 2006. Behavioral ecology and ar-chaeology. Journal of Archaeological Research 14:143–188.

Bray, W. B., L. Herrera, M. Cardale Schrimpff, P. Botero, and J. G.Monsalve. 1987. The ancient agricultural landscape of Calima,Colombia. In Pre-Hispanic agricultural fields in the Andean region.W. M. Denevan, K. Mathewson, and G. Knapp, eds. Pp. 443–481.British Archaeological Reports International 359. Manchester:British Archaeological Reports International.

Brown, T. A., M. K. Jones, W. Powell, and R. G. Allaby. 2009. Thecomplex origins of domesticated crops in the Fertile Crescent.Trends in Ecology & Evolution 24:103–109.

Burbridge, R. E., F. E. Mayle, and T. J. Killeen. 2004. Fifty-thousand-year vegetation and climate history of Noel Kempff Mercado Na-tional Park, Bolivian Amazon. Quaternary Research 61:215–230.

Bush, M. B., M. C. Miller, P. E. DeOlivera, and P. A. Colinvaux.2000. Two histories of environmental change and human distur-bance in eastern lowland Amazonia. Holocene 10:543–553.

Bush, M. B., D. R. Piperno, and P. A. Colinvaux. 1989. A 6,000 yearhistory of Amazonian maize cultivation. Nature 340:303–305.

Carroll, S. B. 2005. Endless forms most beautiful: the new science ofevo devo. New York: Norton.

Castillo, N., and F. J. Aceituno. 2006. El bosque domesticado, elbosque cultivado: un proceso milenario en el valle medio del RıoPorce en el Noroccidente colombiano. Latin American Antiquity17:561–578.

Cavalier, I., C. Rodrıguez, L. F. Herrera, G. Morcote, and S. Mora.1995. No solo de caza vive el hombre: ocupacion del bosque Ama-zonico, holoceno temprano. In Ambito y ocupaciones tempranas dela America tropical. I. Cavalier and S. Mora, eds. Pp. 27–44. Bogota,Colombia: Instituto Colombiano de Antropologıa, Fundacion Er-igaie.

Chandler-Ezell, K., D. M. Pearsall, and J. A. Zeidler. 2006. Root andtuber phytoliths and starch grains document manioc (Manihotesculenta), arrowroot (Maranta arundinacea), and lleren (Calatheasp.) at the Real Alto site, Ecuador. Economic Botany 60:103–120.

Cohen, D. J. 2011. The beginnings of agriculture in China: a mul-tiregional view. Current Anthropology 52(suppl. 4):S273–S293.

Cross, H., R. Lira Saade, and T. J. Motley. 2006. Origin and diver-sification of chayote. In Darwin’s harvest. T. J. Motley, N. Zerega,and H. Cross, eds. Pp. 171–194. New York: Columbia UniversityPress.

Denham, T. 2011. Early agriculture and plant domestication in NewGuinea and Island Southeast Asia. Current Anthropology 52(suppl.4):S379–S395.

Dickau, R. 2010. Comparing subsistence and the transition to ag-riculture in western and central Panama through the integrationof macrobotanical, microbotanical, and paleoenvironmental in-formation. In Integrating zooarchaeology and paleoethnobotany: aconsideration of issues, methods, and cases. A. M. VanDerwarkerand T. M. Peres, eds. Pp. 99–134. New York: Springer.

Dickau, R., A. J. Ranere, and R. G. Cooke. 2007. Starch grain evidencefor the preceramic dispersals of maize and root crops into tropicaldry and humid forests of Panama. Proceedings of the NationalAcademy of Sciences of the USA 104:3651–3656.

Dillehay, T. D., H. H. Eling Jr., and J. Rossen. 2005. Irrigation canalsin the Peruvian Andes. Proceedings of the National Academy ofSciences of the USA 102:17241–17244.

Dillehay, T. D., P. J. Netherly, and J. Rossen. 1989. Middle preceramicpublic and residential sites on the forested slope of the WesternAndes, Northern Peru. American Antiquity 54:733–759.

Dillehay, T. D., C. Ramirez, M. Pino, M. B. Collins, J. Rossen, andJ. D. Pino-Navarro. 2008. Monte Verde: seaweed, food, medicine,and the peopling of South America. Science 320:784–789.

Page 17: 1 - Piperno Neotropicos

S468 Current Anthropology Volume 52, Supplement 4, October 2011

Dillehay, T. D., J. Rossen, T. C. Andres, and D. E. Williams. 2007.Preceramic adoption of peanut, squash, and cotton in northernPeru. Science 316:1890–1893.

Dillehay, T. D., J. Rossen, P. Netherly, and A. Karathanasis. 2010.Early Holocene coca production in northern Peru. Antiquity 84:939–953.

Dobney, K., and G. Larson. 2006. Genetics and animal domestication:new windows on an elusive process. Journal of Zoology 269:261–271.

Doebley, J. 2004. The genetics of maize evolution. Annual Review ofGenetics 38:37–59.

———. 2006. Unfallen grains: how ancient farmers turned weedsinto crops. Science 312:1318–1319.

Erickson, D. L., B. D. Smith, A. C. Clarke, D. H. Sandweiss, and N.Tuross. 2005. An Asian origin for a 10,000 year-old domesticatedplant in the Americas. Proceedings of the National Academy ofSciences of the USA 102:18315–18320.

Flannery, K. V. 1986. The research problem. In Guila Naquitz: archaicforaging and early agriculture in Oaxaca, Mexico. K. V. Flannery,ed. Pp. 3–18. Orlando, FL: Academic Press.

Fuller, D. 2007. Contrasting patterns in crop domestication and do-mestication rates: recent archaeological insights from the OldWorld. Annals of Botany 100:903–924.

Gibson, G., and I. Dworkin. 2004. Uncovering cryptic genetic var-iation. Nature Reviews Genetics 5:681–690.

Gnecco, C., and J. Aceituno. 2006. Early humanized landscapes ofnorthern South America. In Paleoindian archaeology: a hemisphericperspective. J. E. Morrow and C. Gnecco, eds. Pp. 86–104. Gaines-ville: University of Florida Press.

Gremillion, K. J. 2004. Seed processing and the origins of food pro-duction in eastern North America. American Antiquity 69:215–233.

Gremillion, K. J., and D. R. Piperno. 2009a. Human behavioral ecol-ogy, phenotypic (developmental) plasticity, and agricultural ori-gins: insights from the emerging evolutionary synthesis. In Re-thinking the origins of agriculture. M. N. Cohen, ed. Special issue,Current Anthropology 50:615–619.

———. 2009b. Parallel causation does not make a prime mover andhistorical particulars do not make an explanation: response. InRethinking the origins of agriculture. M. N. Cohen, ed. Special issue,Current Anthropology 50:709–711.

Griffith, M. P. 2004. The origins of an important cactus crop, Opuntiaficus-indica (Cactaceae): new molecular evidence. American Journalof Botany 91:1915–1921.

Harlan, J. R. 1971. Agricultural origins: centers and non-centers.Science 174:468–474.

———. 1992. Crops and man. 2nd edition. Madison, WI: AmericanSociety of Agronomy and Crop Science Society of America.

Harris, D. R. 1972. The origins of agriculture in the tropics. AmericanScientist 60:180–193.

———. 1989. An evolutionary continuum of people-plant interac-tion. In Foraging and farming: the evolution of plant exploitation.D. R. Harris and G. Hillman, eds. Pp. 11–26. London: UnwinHyman.

Herrera, L., W. Bray, M. Cardale de Schrimpff, and P. Botero. 1992.Nuevas fechas de radiocarbono para el preceramico en la cordilleraoccidental de Colombia. In Archaeology and environment in LatinAmerica. O. R. Ortiz-Troncoso and T. Van Der Hammen, eds. Pp.145–164. Amsterdam: Universiteit van Amsterdam.

Holst, I., J. E. Moreno, and D. R. Piperno. 2007. Identification ofteosinte, maize, and Tripsacum in Mesoamerica by using pollen,starch grains, and phytoliths. Proceedings of the National Academyof Sciences of the USA 104:17608–17613.

Hughes, C. E., R. Govindarajulu, A. Robertson, D. L. Filer, S. A.Harris, and C. Donovan Bailey. 2007. Serendipitous backyard hy-

bridization and the origin of crops. Proceedings of the NationalAcademy of Sciences of the USA 104:14389–14394.

Iriarte, J. 2007. New perspectives on plant domestication and thedevelopment of agriculture in the New World. In Rethinking ag-riculture. T. Denham, J. Iriarte, and L. Vrydaghs, eds. Pp. 167–188. Walnut Creek, CA: Left Coast.

———. 2009. Narrowing the gap: exploring the diversity of earlyfood-production economies in the Americas. In Rethinking theorigins of agriculture. M. N. Cohen, ed. Special issue, Current An-thropology 50:677–680.

Iriarte, J., I. Holst, O. Marozzi, C. Listopad, E. Alonso, A. Rinder-knecht, and J. Montana. 2004. Evidence for cultivar adoption andemerging complexity during the mid-Holocene in the La PlataBasin, Uruguay. Nature 432:614–617.

Jones, M., and T. Brown. 2007. Selection, cultivation, and repro-ductive isolation. In Rethinking agriculture. T. Denham, J. Iriarte,and L. Vrydaghs, eds. Pp. 36–49. Walnut Creek, CA: Left Coast.

Kaplan, L., and T. F. Lynch. 1999. Phaseolus (Fabaceae) in archae-ology: AMS radiocarbon dates and their significance for pre-Co-lumbian agriculture. Economic Botany 53:262–272.

Kennett, D. J., and B. Winterhalder, eds. 2006. Behavioral ecology andthe transition to agriculture. Berkeley: University of California Press.

Kruglyak, L., and D. L. Stern. 2007. An embarrassment of switches.Science 317:758–759.

Kwak, M., J. A. Kami, and P. Gepts. 2009. The putative Mesoamericandomestication center of Phaseolus vulgaris is located in the Lerma-Santiago Basin of Mexico. Crop Science 49:554–563.

Lauter, N., and J. Doebley. 2002. Genetic variation for phenotypicallyinvariant traits detected in teosinte: implications for the evolutionof novel forms. Genetics 160:333–342.

Leotard, G., A. Duputie, F. Kjellberg, E. J. P. Douzery, C. Debain, J.-J. de Granville, and D. McKey. Forthcoming. Phylogeography andthe origin of cassava: new insights from the northern rim of theAmazon basin. Molecular Phylogenetics and Evolution.

Lupo, K. D. 2007. Evolutionary foraging models in zooarchaeologicalanalysis: recent applications and future challenges. Journal of Ar-chaeological Research 15:143–189.

Matsuoka, Y., Y. Vigouroux, M. M. Goodman, J. Sanchez, E. Buckler,and J. Doebley. 2002. A single domestication for maize shown bymultilocus microsatellite genotyping. Proceedings of the NationalAcademy of Sciences of the USA 99:6080–6084.

Milla, S. R., T. G. Isleib, and H. T. Stalker. 2005. Taxonomic rela-tionships among Arachis sect. Arachis species as revealed by AFLPmarkers. Genome 48:1–11.

Miller, A., and B. Schaal. 2005. Domestication of Mesoamerican cul-tivated fruit tree, Spondias purpurea. Proceedings of the NationalAcademy of Sciences of the USA 102:12801–12806.

Mora, S. 2003. Early inhabitants of the Amazonian tropical rain forest:a study of humans and environmental dynamics. Latin AmericanArchaeology Reports, vol. 3. Pittsburgh: University of Pittsburgh.

Mora, S., L. F. Herrera, I. Cavelier, and C. Rodrıguez. 1991. Cultivars,anthropic soils and stability. Latin American Archaeology Reports,vol. 3. Pittsburgh, PA: University of Pittsburgh.

Motta-Aldana, J. R., M. L. Serrano-Serrano, J. H. Torres, G. Castillo-Villamizar, D. G. Debouck, and M. I. Chacon. Forthcoming. Do-mestication patterns in wild lima beans (Phaseolus lunatus L.) fromthe Americas. In Harlan II: biodiversity in agriculture: domestication,evolution, and sustainability. A. Damania and P. Gepts, eds. Cam-bridge: Cambridge University Press.

New, J., M. M. Krasnow, D. Truxaw, and S. J. C. Gaulin. 2007. Spatialadaptations for foraging: women excel and calories count. Pro-ceedings of the Royal Society B 274:2679–2684.

Olsen, K. M. 2002. Population history of Manihot esculenta (Eu-phorbiaceae) inferred from nuclear DNA sequences. MolecularEcology 11:901–911.

Olsen, K. M., and B. Schaal. 1999. Evidence on the origin of cassava:

Page 18: 1 - Piperno Neotropicos

Piperno Plant Cultivation and Domestication in the New World Tropics S469

phylogeography of Manihot esculenta. Proceedings of the NationalAcademy of Sciences of the USA 96:5586–5591.

Pearsall, D. M. 1978. Phytolith analysis of archaeological soils: evi-dence for maize cultivation in Formative Ecuador. Science 199:177–178.

———. 2000. Paleoethnobotany: a handbook of procedures. 2nd edi-tion. San Diego, CA: Academic Press.

———. 2003. Plant food resources of the Ecuadorian Formative: anoverview and comparison to the Central Andes. In Archaeology ofFormative Ecuador. J. S. Raymond and R. L. Burger, eds. Pp. 213–257. Washington, DC: Dumbarton Oaks.

Pearsall, D. M., K. Chandler-Ezell, and A. Chandler-Ezell. 2003. Iden-tifying maize in Neotropical sediments and soils using cob phy-toliths. Journal of Archaeological Science 30:611–627.

Pearsall, D. M., K. Chandler-Ezell, and J. A. Zeidler. 2004. Maize inancient Ecuador: results of residue analysis of stone tools from theReal Alto site. Journal of Archaeological Science 31:423–442.

Perry, L., R. Dickau, S. Zarillo, I. Holst, D. M. Pearsall, D. R. Piperno,M. J. Berman, et al. 2007. Starch fossils and the domestication anddispersal of chile peppers (Capsicum sp.) in the Americas. Science315:986–988.

Pigliucci, M. 2005. Expanding evolution. Nature 435:565–566.Piperno, D. R. 1988. Phytolith analysis: an archaeological and geological

perspective. San Diego, CA: Academic Press.———. 1990. Aboriginal agriculture and land usage in the Amazon

Basin, Ecuador. Journal of Archaeological Science 17:665–677.———. 2006a. A behavioral ecological approach to the origins of

plant cultivation and domestication in the seasonal tropical forestsof the New World. In Foraging theory and the transition to agri-culture. D. Kennett and B. Winterhalder, eds. Pp. 137–166. Berke-ley: University of California Press.

———. 2006b. Identifying manioc (Manihot esculenta Crantz) andother crops in pre-Columbian tropical America through starchgrain analysis: a case study from Panama. In Documenting do-mestication: new genetic and archaeological paradigms. M. Zeder,E. Emschwiller, D. Bradley, and B. S. Smith, eds. Pp. 46–67. Berke-ley: University of California Press.

———. 2006c. Phytoliths: a comprehensive guide for archaeologistsand paleoecologists. Lanham, MD: AltaMira.

———. 2009. Identifying crop plants with phytoliths (and starchgrains) in Central and South America: a review and an update ofthe evidence. Quaternary International 193:146–159.

———. Forthcoming. New archaeobotanical information on earlycultivation and plant domestication involving microplant remains.In Harlan II: biodiversity in agriculture: domestication, evolution,and sustainability. A. Damania and P. Gepts, eds. Cambridge: Cam-bridge University Press.

Piperno, D. R., K. Clary, R. G. Cooke, A. J. Ranere, and D. Weiland.1985. Preceramic maize in central Panama: phytolith and pollenevidence. American Anthropologist 87:871–878.

Piperno, D. R., and T. D. Dillehay. 2008. Starch grains on humanteeth reveal early broad crop diet in northern Peru. Proceedings ofthe National Academy of Sciences of the USA 105:19622–19627.

———. Forthcoming. Agricultural origins and their social impli-cations in South America. In The Cambridge prehistory. C. Renfrewand P. Bahn, eds. Cambridge: Cambridge University Press.

Piperno, D. R., J. E. Moreno, I. Holst, M. Lachniet, J. G. Jones, A.J. Ranere, and R. Castanzo. 2007. Late Pleistocene and Holoceneenvironmental history of the Iguala Valley, central Balsas watershedof Mexico. Proceedings of the National Academy of Sciences of theUSA 104:11874–11881.

Piperno, D. R., and D. M. Pearsall. 1998. The origins of agriculturein the lowland Neotropics. San Diego, CA: Academic Press.

Piperno, D. R., A. J. Ranere, I. Holst, and P. Hansell. 2000. Starchgrains reveal early root crop horticulture in the Panamanian trop-ical forest. Nature 407:894–897.

Piperno, D. R., A. J. Ranere, I. Holst, J. Iriarte, and R. Dickau. 2009.Starch grain and phytolith evidence for early ninth millenniumBP maize from the Central Balsas River Valley, Mexico. Proceedingsof the National Academy of Sciences of the USA 106:5019–5024.

Piperno, D. R., and K. Stothert. 2003. Phytolith evidence for earlyHolocene Cucurbita domestication in southwest Ecuador. Science299:1054–1057.

Plowman, T. 1984. The origin, evolution, and diffusion of coca, Er-ythroxylum sp., in South and Central America. In Pre-Columbianplant migration. D. Stone, ed. Pp. 126–163. Cambridge, MA: Pea-body Museum, Harvard University.

Pohl, M. E. D., D. R. Piperno, K. O. Pope, and J. G. Jones. 2007.Microfossil evidence for pre-Columbian maize dispersals in theNeotropics from San Andres, Tabasco, Mexico. Proceedings of theNational Academy of Sciences of the USA 104:6870–6875.

Pope, K. O., M. E. D. Pohl, J. G. Jones, D. L. Lentz, C. Von Nagy,F. J. Vega, and I. R. Quitmyer. 2001. Origin and environmentalsetting of ancient agriculture in the lowlands of Mesoamerica.Science 292:1370–1373.

Ranere, A. J., and R. G. Cooke. 2003. Late glacial and early Holoceneoccupation of Central American tropical forests. In Under the can-opy: the archaeology of tropical rain forests. J. Mercader, ed. Pp.219–248. Piscataway, NJ: Rutgers University Press.

Ranere, A. J., D. R. Piperno, I. Holst, R. Dickau, and J. Iriarte. 2009.Preceramic human occupation of the Central Balsas Valley, Mexico:cultural context of early domesticated maize and squash. Proceed-ings of the National Academy of Sciences of the USA 106:5014–5018.

Raymond, J. S. 2008. The process of sedentism in northwestern SouthAmerica. In Handbook of South American archaeology. H. Silver-man and W. H. Isbell, eds. Pp. 79–90. New York: Springer.

Richardson, J. B., III. 1973. The preceramic sequence and the Pleis-tocene and post-Pleistocene climate of northwest Peru. In Vari-ations in anthropology: essays in honor of John C. McGregor. D. W.Lathrap and J. Douglas, eds. Pp. 199–211. Urbana: Illinois Ar-chaeological Survey.

Rindos, D. 1984. The origins of agriculture: an evolutionary perspective.Orlando, FL: Academic Press.

Robledo, G., G. I. Lavia, and G. Seijo. 2009. Species relations amongwild Arachis species with the A genome as revealed by FISH map-ping of rDNA loci and heterochromatin detection. Theoretical andApplied Genetics 118:1295–1307.

Rodrigues, D. P., S. A. Filho, and C. R. Clement. 2004. Molecularmarker-mediated validation of morphologically defined landracesof Pejibaye (Bactris gasipaes) and their phylogenetic relationships.Genetic Resources and Crop Evolution 51:871–882.

Rosas, R. V., T. Andres, and M. Nee. 2004. The Goldman cucurbitcollecting expedition in Peru, 2004. http://www.cucurbit.org/Peru/PeruReport.html.

Salgado, H. 1995. El preceramico en el canon del rıo Calima, cor-dillera occidental. In Ambito y ocupaciones tempranas de la AmericaTropical. I. Cavelier and S. Mora, eds. Pp. 73–82. Bogota: Fun-dacion Erigaie, Instituto Colombiano de Antropologıa.

Sandweiss, D. H. 2003. Terminal Pleistocene through mid-Holocenearchaeological sites as paleoclimatic archives for the Peruvian coast.Palaeogeography, Palaeoclimatology, Palaeoecology 194:23–40.

Sanjur, O., D. R. Piperno, T. C. Andres, and L. Wessell-Beaver. 2002.Phylogenetic relationships among domesticated and wild speciesof Cucurbita (Cucurbitaceae) inferred from a mitochondrial gene:implications for crop plant evolution and areas of origin. Pro-ceedings of the National Academy of Sciences of the USA 99:535–540.

Sauer, C. O. 1950. Cultivated plants of South and Central America.In Handbook of South American Indians. J. Steward, ed. Pp. 487–543. Bureau of American Ethnology, Bulletin 143, vol. 6. Wash-ington, DC: U.S. Government Printing Office.

Shady Solis, R., J. Haas, and W. Creamer. 2001. Dating Caral, a

Page 19: 1 - Piperno Neotropicos

S470 Current Anthropology Volume 52, Supplement 4, October 2011

preceramic site in the Supe Valley on the central coast of Peru.Science 292:723–726.

Shennan, S. 2008. Evolution in archaeology. Annual Review of An-thropology 37:75–91.

Simms, S., and K. W. Russell. 1997. Bedouin hand harvesting ofwheat and barley: implications for early cultivation in South-western Asia. Current Anthropology 38:696–702.

Smith, B. D. 2009. Core conceptual flaws in human behavioral ecol-ogy. Communicative and Integrative Biology 2:533–534.

Stothert, K. 1985. The preceramic Las Vegas culture of coastal Ec-uador. American Antiquity 50:613–637.

Stothert, K., D. R. Piperno, and T. C. Andres. 2003. Terminal Pleis-tocene/early Holocene human adaptation in coastal Ecuador: theLas Vegas evidence. Quaternary International 109–110:23–43.

van Heerwaarden, J., J. Doebley, W. H. Briggs, J. C. Glaubitz, M. M.Goodman, J. J. Sanchez Gonzalez, J. Ross-Ibarra, B. S. Gaut, andL. Eguiarte. 2011. Genetic signals of origin, spread, and intro-gression in a large sample of maize landraces. Proceedings of theNational Academy of Sciences of the USA 108:1088–1092.

Vrydaghs, L., and T. Denham. 2007. Rethinking agriculture: intro-ductory thoughts. In Rethinking agriculture. T. Denham, J. Iriarte,and L. Vrydaghs, eds. Pp. 1–15. Walnut Creek, CA: Left Coast.

Weir, G. H., and J. P. Dering. 1986. The lomas of Paloma: human-environment relations in a central Peruvian fog oasis: archaeo-botany and palynology. In Andean archaeology: papers in memoryof Clifford Evans. R. Matos, S. A. Turpin, and H. H. Eling Jr., eds.Pp. 18–44. Los Angeles: Institute of Archaeology, University ofCalifornia.

Weiss, E., M. E. Kislev, and A. Hartmann. 2006. Autonomous cul-tivation before domestication. Science 312:1608–1610.

Weiss, E., and D. Zohary. 2011. The Neolithic Southwest Asian foun-der crops: their biology and archaeobotany. Current Anthropology52(suppl. 4):S237–S254.

Wessel-Beaver, L. 2000. Evidence for the center of diversity of Cu-curbita moschata in Colombia. Cucurbit Genetics Cooperative Re-port 23:54–55.

West-Eberhard, M. J. 2003. Developmental plasticity and evolution.Oxford: Oxford University Press.

Westengen, O. T., Z. Huaman, and M. Heun. 2005. Genetic diversityand geographic pattern in early South American cotton domes-tication. Theoretical and Applied Genetics 110:392–402.

Willcox, G., S. Fornite, and L. Herveux. 2008. Early Holocene cul-tivation before domestication in northern Syria. Vegetational His-tory and Archaeobotany 17:313–325.

Zarillo, S., D. M. Pearsall, J. S. Raymond, M. A. Tisdale, and D. J.Quon. 2008. Directly dated starch residues document Early For-mative maize (Zea mays L.) in tropical Ecuador. Proceedings of theNational Academy of Sciences of the USA 105:5005.

Zeder, M. A. 2011. The origins of agriculture in the Near East. CurrentAnthropology 52(suppl. 4):S221–S235.

Zeder, M. A., and B. D. Smith. 2009. A conversation on agriculturalorigins: talking past each other in a crowded room. Current An-thropology 50:681–691.

Zhang, D. P., J. Cervantes, Z. Huaman, E. Carey, and M. Ghislain.2000. Assessing genetic diversity of sweet potato (Ipomoea batatas(L.) Lam.) using cultivars from tropical America using AFLP. Ge-netic Resources and Crop Evolution 47:659–665.