Hostname: page-component-76fb5796d-vvkck Total loading time: 0 Render date: 2024-04-26T06:33:10.765Z Has data issue: false hasContentIssue false

THEORETICAL AND SOCIOECOLOGICAL CONSEQUENCES OF FIRE FOODWAYS

Published online by Cambridge University Press:  16 October 2018

Alan P. Sullivan III*
Affiliation:
Department of Anthropology, University of Cincinnati, 481 Braunstein Hall, P.O. Box 210380, Cincinnati, OH 45221-0380, USA
Philip B. Mink II
Affiliation:
W. S. Webb Museum of Anthropology, University of Kentucky, Lexington, KY 40506-0027, USA
*
(alan.sullivan@uc.edu, corresponding author)
Rights & Permissions [Opens in a new window]

Abstract

Archaeological investigations of the effects of anthropogenic fire on the subsistence economies of small-scale societies, particularly those of the prehispanic northern American Southwest, are embryonic in scope and disciplinary impact. When burning has been mentioned in such studies it typically has been with reference to its alleged effectiveness in clearing land or deforesting areas for maize agriculture. In this article, in contrast, we present the results of our initial efforts to estimate the yield and socioecological consequences of cultivating a common fire-responsive ruderal—amaranth—whose growth is enabled by anthropogenic burning of understory vegetation in the Southwest's pinyon-juniper ecosystems. With data from the Upper Basin (northern Arizona), we show that, in an area that is not environmentally conducive to maize production, populations could be supported with systematic, low-intensity anthropogenic fires that promoted the growth of amaranth and other ruderals, such as chenopodium, which consistently dominate archaeobotanical and pollen assemblages recovered from a variety of archaeological and sedimentary contexts in the region. Based on this evidence, as well as modern fire ecological data, we propose that fire-reliant ruderal agriculture, in contrast to maize agriculture, was a widespread, sustainable, and ecologically sound practice that enhanced food supply security independently of variation in soil fertility and precipitation.

Las investigaciones arqueológicas sobre los efectos de los incendios antropogénicos para las economías de subsistencia de las sociedades de pequeña escala, especialmente aquellas de la zona norte del suroeste norteamericano en la época precolombina, se encuentran todavía en un estado naciente y tienen poca influencia en la disciplina. Cuando se mencionan los incendios en tales estudios, es típicamente en referencia a su supuesta eficacia para el desmonte o la deforestación de tierras antes de sembrar maíz. En contraste, en este artículo presentamos la primera estimación del rendimiento y de las consecuencias socio-ecológicas del cultivo de amaranto, una especie ruderal común cuyo crecimiento incrementa en respuesta al incendio antropogénico de la vegetación del sotobosque en el ecosistema piñón-junípero del suroeste norteamericano. Con datos procedentes de la cuenca superior del Río Colorado, en el norte de Arizona, demostramos que en áreas marginales para la cultivación del maíz las comunidades agrícolas pudieron causar incendios de baja intensidad para promover el crecimiento del amaranto y otros ruderales tales como el quenopodio —plantas que dominan las muestras de polen arqueológico en sedimentos encontrados en esta región. Con base en esta evidencia y en datos recientes sobre la ecología del fuego, planteamos que una agricultura ruderal dependiente de los incendios, en contraste con el cultivo del maíz, fue generalizada, sostenible, ecológicamente saludable, e incrementó la seguridad de la provisión de alimentos independientemente de variaciones en la fertilidad de la tierra y precipitación.

Type
Articles
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
Copyright © 2018 by the Society for American Archaeology

Cross-disciplinary understanding of the transformative effects of anthropogenic landscape fire on ecosystems, their structure, and associated “services” has accelerated dramatically in recent decades (e.g., Bowman et al. Reference Bowman, Balch, Artaxo, Bond, Carlson, Cochrane, D'Antonio, DeFries, Doyle, Harrison, Johnston, Keeley, Krawchuk, Kull, Brad Marston, Moritz, Colin Prentice, Roos, Scott, Swetnam, van der Werf and Pyne2009). Thinking globally about these developments in the context of human prehistory, it is rare for modern research regarding human-environment interactions not to mention anthropogenic fire as one of the principal ecosystem-shaping forces during the Pleistocene and Holocene (Supplemental Text 1). Archaeological studies worldwide, ranging from the Upper Paleolithic (Haws Reference Haws2012:72) and Mesolithic (Mason Reference Mason2000) in Western Europe, to the early Neolithic in Southeast Asia (Hunt and Rabett Reference Hunt and Rabett2014:26), to the late Holocene in eastern North America (Abrams and Nowacki Reference Abrams and Nowacki2008), indicate that human-controlled fire influenced the nature of social formations and food supply systems of the ancient and modern worlds (Bond and Keeley Reference Bond and Keeley2005). These investigations are especially timely in view of the current attention directed at the contributions of fire-induced particulates to climate change (Han et al. Reference Han, Peteet, Arimoto, Cao, An, Sritrairat and Yan2016) and the continuing controversy over whether the onset and duration of the Anthropocene (Braje Reference Braje2015) should be defined in terms of atmospheric chemistry (Ruddiman Reference Ruddiman2013) or domestication processes (Smith and Zeder Reference Smith and Zeder2013).

Interestingly, these studies are unified by a topic with deep historical roots and broad interdisciplinary connections—understanding the relation between anthropogenic landscape fire ecology and subsistence economies (Supplemental Text 2). As many environmental historians have remarked, appreciating the ecological dynamics and evolutionary consequences of this entangled relationship intrinsically engages archaeology (e.g., Bonnicksen Reference Bonnicksen2000). However, archaeological data that reflect the extent to which humans persistently employed landscape fire for subsistence purposes are “surprisingly scarce” (Scherjon et al. Reference Scherjon, Berkels, MacDonald and Roebroeks2015:321). One complicating factor is that, despite the rich ethnographic, ethnohistoric, and historic accounts of humans igniting landscape fires for a variety of reasons (Supplemental Text 3), including wild plant husbandry, game management, and pest control (Huffman Reference Huffman2013; Roos Reference Roos, Mills and Fowles2017), such descriptions provide few details that archaeologists can draw upon to inform their investigations of anthropogenic fire and its economic aftermath (Lightfoot et al. Reference Lightfoot, Cuthrell, Striplen and Hylkema2013:286).

One approach that has gained considerable attention in the American Southwest, however, is applied historical ecology (Swetnam et al. Reference Swetnam, Allen and Betancourt1999), which infers the effects of landscape burning by examining “paleofire proxies,” such as fire scar, sedimentologic, palynological, and geoanthracological (detrital charcoal) records, and carbon isotope ratios of soil organic matter (French et al. Reference French, Perriman, Cummings, Hall, Goodman-Elgar and Boreham2009; Roos Reference Roos, Scheiber and Zedeño2015). The results of these theoretically robust and empirically rich studies demonstrate that anthropogenic landscape burning was indeed a transformative landscape management and ecosystem-structuring technique (Liebmann et al. Reference Liebmann, Farella, Roos, Stack, Martini and Swetnam2016; Roos and Swetnam Reference Roos and Swetnam2012). Nevertheless, the applicability of applied historical ecology is constrained by the “fading record” problem (e.g., most fire scar records in the American Southwest postdate AD 1500; Fulé et al. Reference Fulé, Heinlein, Wallace Covington and Moore2003) and by the likelihood that knowledge of the range of fire regimes may be historically biased because of the “no analogue” problem (Swetnam et al. Reference Swetnam, Allen and Betancourt1999:1192, 1198).

We see these issues, however, as opportunities to expand the usefulness of applied historical ecology by exploring the possibility that, in regions prone to the record-fading and no-analogue problems, such as the Coconino Plateau and Grand Canyon (Williams and Baker Reference Williams and Baker2013:298), the effects of low-intensity anthropogenic fire would register in archaeobotanical remains recovered from well-dated archaeological contexts (Miller and Tausch Reference Miller, Tausch, Galley and Wilson2001:17). Specifically, we posit that people intentionally burned understory vegetation in Southwestern pinyon-juniper woodlands to produce fire-responsive ruderals (Sullivan Reference Sullivan, Ingram and Hunt2015) that, once harvested, processed, and discarded in a variety of locations, became concentrated in and around now-abandoned settlements (Yarnell Reference Yarnell1965). This proposition underscores the centrality of incorporating direct evidence from the archaeobotanical record in establishing the role of fire in economic prehistory (Smith Reference Smith, Marston, Guedes and Warinner2014:369), particularly for those periods of occupation in the American Southwest for which there is scant ethnographic documentation for fire-related subsistence practices (Roos Reference Roos, Mills and Fowles2017; Sullivan and Forste Reference Sullivan and Forste2014).

Our objective here is to illustrate the explanatory potential of the fire foodway model for the American Southwest's vast pinyon-juniper woodlands,Footnote 1 which, in contrast to maize (Zea mays) agriculture, takes advantage of the principles of fire ecology and aligns with the existing archaeobotanical record (Ford Reference Ford1981:6). We first discuss the key elements of current maize agriculture farming models for pinyon-juniper woodlands, which intrinsically do not consider burning a cultivation method (except for land clearing; Crabtree et al. Reference Crabtree, Vaughn and Crabtree2017:125; Wyckoff Reference Wyckoff1977). Next, adopting the basic structure of maize farming and productivity models (Bocinsky and Varien Reference Bocinsky and Varien2017; Kohler Reference Kohler, Kohler and Varien2012), we determine the per capita caloric intake needed to sustain one individual on only one ruderal species (amaranth; Amaranthus spp.), derive population estimates for one human generation (25 years), and develop productivity estimates for amaranth to support various levels of population during different periods of occupation in the Upper Basin. Then, we introduce the fire foodway model of ruderal cultivation, which is based on two well-secured ethnographic findings: (1) creating anthropogenic niches with burning is a common ecosystem-transforming technique worldwide (Smith Reference Smith and Smith2011), and (2) burn plots established within anthropogenic niches, whose use is rotated on a two- to three-year cycle, take advantage of the invariable appearance of ruderals during the earliest postburn successional cycle (Everett and Ward Reference Everett and Ward1984). Finally, inspired by firsthand contemporary observations of how ruderals predictably respond to a variety of fire types, severity, and origin, we offer some thoughts about how these considerations have the potential to integrate applied historical ecology and niche construction theory to enrich our narratives of past human-environment interactions and economic prehistory in the American Southwest.

Maize Farming in Pinyon-Juniper Woodlands

Maize productivity modeling studies for pinyon-juniper woodlands in the upland Southwest are based on the following propositions:

Discussion

The fire foodway model does not rely on any of these assumptions but is informed, instead, by two understandings. First, the production of fire-responsive (or fire-stimulated [Nabhan et al. Reference Nabhan, Coder and Smith2004:18–19]) economic plants, such as amaranth, chenopodium (Chenopodium spp.), and various grasses (Bohrer Reference Bohrer1975), promotes a secure livelihood in conditions that are considered marginal for maize farming (e.g., Benson et al. Reference Benson, Ramsey, Stahle and Petersen2013; Sullivan Reference Sullivan, Tainter and Tainter1996). These ruderals, which typically colonize and thrive in human-created disturbances or niches (Smith Reference Smith, Marston, Guedes and Warinner2014), have been (1) categorized as weeds or inadvertent by-products of maize farming (e.g., Ford Reference Ford, Plog and Powell1984), (2) designated as starvation or famine foods (Minnis Reference Minnis1991), or (3) asserted to have been rarely (if ever) cultivated or the focus of sustained cultivation (e.g., Plog et al. Reference Plog, Fish, Glowacki and Fish2015:11) despite widespread archaeological evidence to the contrary (Table 1; Fritz Reference Fritz and Twiss2007:289–291). In fact, the single plant we focus on here, amaranth, has a history of cultivation and use that is as long as that of maize (Supplemental Text 4), and its yield “per unit of land may be greater than that of corn” (Jones Reference Jones1953:91; see also Bohrer Reference Bohrer and Caywood1962:108). Ethnographic (Anscheutz Reference Anscheutz, Price and Morrow2006) and ethnobotanical (Bye Reference Bye1981) studies likewise attest to the economic significance of amaranth, as well as chenopodium, in ancient and modern Southwestern subsistence economies (Doolittle Reference Doolittle2000:95; Ford Reference Ford, Minnis and Elisens2000:217; Morrow Reference Morrow, Price and Morrow2006:22). Second, Southwest archaeologists have rarely thought of fire itself as an applied technology for direct food production (e.g., Adams and Fish Reference Adams, Fish and Smith2011:167–170), even though such applications are common worldwide (Dods Reference Dods2002; Roos et al. Reference Roos, Scott, Belcher, Chaloner, Aylen, Bird, Coughlan, Johnson, Johnston, McMorrow and Steelman2016:4–5).

Table 1. Archaeological Studies that Mention Amaranth, Chenopodium, or “Cheno-ams” as Cultivated or Economically Significant Plants in the American Southwest.

Note: Studies presented in this table were selected to maximize spatial and temporal variability and date of publication.

Suitability of the Upper Basin for Maize-Based Foodways

Our study area is the Upper Basin, which is a downfaulted and tilted graben of the Coconino Plateau that extends south of the eastern South Rim of the Grand Canyon (2,256 m asl at Desert View) to the base of the Coconino Rim (1,859 m asl at Lee Canyon; Figure 1). Today, the Upper Basin is blanketed by a dense pinyon-juniper woodland (Vankat Reference Vankat2013) that becomes intermixed with ponderosa pine on its western edge but grades to grassland farther south (Darling Reference Darling1967). Like so many areas in the upland Southwest occupied between AD 875 and 1200 (Euler Reference Euler and Gumerman1988), the Upper Basin is thickly stocked with abandoned one- to two-room structures and other features, such as rock alignments and terraces (Sullivan et al. Reference Sullivan, Berkebile, Forste and Washam2015), which conventionally have been interpreted as landscape signatures of maize production (e.g., Effland et al. Reference Effland, Jones and Euler1981; Stewart and Donnelly Reference Stewart and Donnelly1943). However, these appearances are deceiving when we examine the area's modern and ancient environmental characteristics and its archaeo-economic record and evaluate the extent to which they align with the attributes of maize-based foodways.

Figure 1. Location of the Upper Basin, northern Arizona, showing excavated sites and area burned by the Scott Fire.

Soils

Pinyon-juniper woodlands are notorious for establishing themselves on nutrient-poor soils (West Reference West, Anderson, Fralish and Baskin1999:289), and the Upper Basin is no exception. Surface material ranges from bedrock and very cobbly/very gravelly loams in the Upper Basin's upper and central reaches (considered “agriculturally unsuitable” according to the Natural Resources Conservation Service; Lindsay et al. Reference Lindsay, Strait and Denny2003; Merkle Reference Merkle1952:377) to very gravelly/gravelly sandy loams in its lower reaches (suitable only for rangeland after “conversion;” Figure 2; Brewer et al. Reference Brewer, Jorgensen, Munk, Robbie and Travis1991). Soil chemical and texture analyses of archaeological terrace sediments (Homburg Reference Homburg and Whittlesey1992; Sullivan Reference Sullivan2000) attest to the Upper Basin's thin and rocky soils (Hendricks Reference Hendricks1985). Pollen analyses of these anthropogenic terrace sediments yielded only a dozen or so Zea mays pollen grains (<0.5%) out of thousands examined, and two sets of samples contained no maize pollen whatsoever (Bozarth Reference Bozarth and Whittlesey1992; Davis Reference Davis and Sullivan1986).

Figure 2. Agriculturally unproductive soils in the Upper Basin: (a) bedrock and very cobbly loam; (b) cobbly and very gravelly sandy loam; (c) very gravelly/sandy loam. (Color online)

Precipitation

With respect to the other major constraint on maize production—water availability during the frost-free growing season—hydrologic studies in the Upper Basin confirm that runoff from rainfall or snowmelt is negligible (Rand Reference Rand1965:13–14), which means that water cannot be directed to where it might be needed for maize production (Metzger Reference Metzger1961; cf. Benson Reference Benson2011:40–41). Further, interannual variation in the Upper Basin's paleoprecipitation patterns was so unpredictable (Sullivan and Ruter Reference Sullivan, Ruter, Doyel and Dean2006:185–188)Footnote 2 that, given the sensitivity of maize to the timing and amount of rain it needs to germinate (Adams Reference Adams, Ingram and Hunt2015:29–32), successful harvests were undoubtedly uncommon events (Schwartz et al. Reference Schwartz, Kepp and Chapman1981:121; see also Bocinsky and Varien Reference Bocinsky and Varien2017:299). On the other hand, importantly, even the lowest values of tree ring–reconstructed annual precipitation for the Upper Basin, that is, 25.4 cm in AD 1067, are more than adequate to ensure the survival of ruderals, such as amaranth (Salt Spring Seeds 2014:3) or chenopodium (Chenopodium quinoa [Smith Reference Smith2017:2]).

Discussion

In terms of soil fertility and frost-free precipitation, the two principal factors stipulated in maize productivity modeling studies, the Upper Basin's pinyon-juniper woodland is a “hostile” environment that is ill-suited for maize production, even during the best of times (i.e., when annual precipitation equaled or exceeded the average of 36.6 cm). It is hardly surprising, therefore, to learn that no ethnographic, ethnohistoric, or historic accounts mention maize farming in the Upper Basin or on the eastern Coconino Plateau by Native Americans or Euro-Americans (Begay and Roberts Reference Begay, Roberts and Towner1996:199–202; Cleeland et al. Reference Cleeland, Hanson, Lesko and Weintraub1990; Hough and Brennan Reference Hough, Brennan and Berger2008; Martin Reference Martin1985). These factors explain, as well, why we have encountered negligible amounts of maize remains such as botanicals or pollen (for comparable results from Black Mesa, see Ford Reference Ford, Plog and Powell1984:131; Ruppé Reference Ruppé, Christenson and Perry1985:521) in archaeological contexts that typically are associated with maize farming in the upland Southwest (Sullivan and Forste Reference Sullivan and Forste2014). In fact, no matter which consumption or production contexts are examined—structures or processing areas (Supplemental Table 1)—the archaeo-economic assemblages from them are overwhelmingly and consistently dominated by ruderal seeds and pinyon nuts (Figure 3).

Figure 3. Variation in the abundance and ubiquity of seeds and nuts (n = 3,485) identified in 110 samples recovered from features (postholes, thermal features, fire-cracked rock piles), artifacts (vessels, grinding stones), and occupation surfaces at 10 archaeological sites in the Upper Basin and Grand Canyon National Park (see Supplemental Table 1 for details). Each dot represents the frequency of nuts or seeds in a single sample, broken down by taxon (seeds classified by different archaeobotanists as cheno-am seeds, amaranth seeds, or chenopodium seeds are aggregated as “Cheno-am”). The data do not include counts of indirect indicators of plant use, such as cone scales, seed coats, nutshell, bark, needles, stems, leaves, wood, or cupules. This method was selected because it tightly constrains frequencies of edible plant parts—seeds or nuts—that in all likelihood were the objects of wild plant cultivation, wild plant gathering, or domesticated plant cultivation (Sullivan et al. Reference Sullivan, Berkebile, Forste and Washam2015:44). Ubiquity values are given in parentheses.

Suitability of the Upper Basin for Fire-Based Ruderal Foodways

Estimating Amaranth Consumption and Production

To illustrate the feasibility of the fire foodway provisioning strategy, we estimate how much amaranth would be required to support one person for a year. This estimate is based on the same parameters featured in numerous maize-based productivity studies and, therefore, involves the following considerations (based on Kohler et al. Reference Kohler, Kresl, West, Carr, Wilshusen, Kohler and Gumerman2000:160; Pool Reference Pool, Wingard and Hayes2013:96; Van West and Lipe Reference Van West, Lipe and Lipe1992:111):

  • One person requires at least 2,000 kcal/day, which is roughly the midpoint in the published range for preindustrial populations (1,560 to 2,550 kcal/day).

  • Sixty-nine percent of the annual diet was amaranth, which is roughly the midpoint of the published range for maize consumption (60% to 77%).

  • One kilogram of amaranth yields 3.7 kcal (Putnam et al. Reference Putnam, Oplinger, Doll and Schulte1989:4), which is comparable to maize (3.5 kcal/kg).

With these understandings, one person would require 0.37 kg of amaranth per day, or 136 kg per year.

Although estimates for amaranth (and, for comparison, chenopodium) productivity vary widely, ranging from 340.25 to 4,310 kg/ha (Supplemental Table 2), we chose the lowest yield value (Amaranth 2) because it is from modern hand-cultivated amaranth farming, which we think more closely approximates prehistoric cultivation practices. Adoption of this value means that 0.40 ha (ca. 1 ac) of land would be needed to produce enough amaranth to satisfy one person's needs per year.

Estimating Population

To estimate the number of people that amaranth could have supported between AD 875 and 1200 in the Upper Basin, we projected the number of room spaces (n = 886) that surveys have recorded (Foust Reference Foust2015) and “time-corrected” it to yield the number of architectural spaces per 25-year period (Downum and Sullivan Reference Downum, Sullivan and Anderson1990), or roughly one human generation (cf. Roberts Reference Roberts, Miller, Moore and Ryan2011:13). These “corrections” allow us to maximize population in the face of Grand Canyon's complicated and incompletely understood population history, which involves several groups (Cohonina, Virgin Anasazi [sensu Euler Reference Euler, Euler and Tikalsky1992], and Grand Canyon Anasazi [sensu Schwartz Reference Schwartz1990]) whose perennial settlements were occupied no longer than 10–15 years (Mink Reference Mink2015; cf. Matson et al. Reference Matson, Lipe and Haase1988:253–254; Ortman et al. Reference Ortman, Diederichs, Schleher, Fetterman, Espinosa and Sommer2016:235).

Next, we calibrated the time-corrected room counts by a perennial occupancy rate. This calculation involved dividing the number of structures (n = 324) whose artifact density exceeded seven artifacts/m2 (based on an excavated perennial settlement [Sullivan Reference Sullivan, Holdaway and Wandsnider2008]) by the number of room spaces (n = 604) for which we have controlled, standardized artifact inventories (Uphus Reference Uphus2003). To convert time-corrected and perennially occupied room counts into numbers of people, we multiplied those values by 2.5 (which is the average interior floor area of 226 single-room structures [13.8 m2] divided by 5.6 m2/person [Liebmann et al. Reference Liebmann, Farella, Roos, Stack, Martini and Swetnam2016]). These estimates overrepresent the maximum number of people to be fed during any year (Supplemental Text 5). Finally, these figures were multiplied by 0.40 ha to yield the maximum number of hectares to be cultivated per year to sustain the estimated population on amaranth alone (Figure 4).

Figure 4. Estimated time-corrected room counts, annual population, and hectares to be cultivated under different yields (kg/ha) of amaranth and chenopodium (based on data in Supplemental Table 2).

Managing Fire-Foodway Ecological Impacts on the Landscape

In modeling the ecological impact of fire-based amaranth production, our assumption is that the Upper Basin's occupants ignited many small fires rather than a few large conflagrations (Adams Reference Adams, Nelson and Strawhacker2011:125; Scherjon et al. Reference Scherjon, Berkels, MacDonald and Roebroeks2015:309). To constrain these estimates further, we rely on Mellars's (Reference Mellars1976) classic article in which he posits that, for pinyon-juniper woodlands in northern Arizona, a burn plot no larger than 400 m in diameter (12.6 ha) maximizes ruderal production. Applying this figure to the quantities of amaranth needed to support varying numbers of people, based on differing productivity estimates, means that between five and 72 400 m diameter fires would have to be ignited per year (Table 2), burning at most no more than 4.2% of the Upper Basin (218.73 km2, excluding unburnable locations, such as bedrock and rock-filled drainages).

Table 2. Estimated Number of Fires Needed per Year Based on Different Productivity Estimates for Amaranth and Chenopodium.

Note: Estimated fires are 400 m in diameter. See Supplemental Table 2 for background data.

aPeriod 1 = AD 875–1000; Period 2 = AD 1000–1050; Period 3 = AD 1050–1075; Period 4 = AD 1075–1115; Period 5 = AD 1115–1200.

bValue in parentheses indicates productivity estimate in hectares per person per year (ha/person/year).

Combining the facultative ecological orientation of niche construction theory with the methodological robustness of applied historical ecology, we propose that anthropogenic niches 1.1 km in diameter (0.95 km2) were established to enable the ignition of 400 m diameter burn plots within which ruderal production occurred (Figure 5; this configuration is comparable conceptually to the 200 m2 [4 ha] cells that have been used for maize production modeling in southwest Colorado [e.g., Kohler and Van West Reference Kohler, Van West, Tainter and Tainter1996:179]). Furthermore, we suggest that to ensure sufficient fuel accumulation to carry low-intensity surface fires, to avoid nutrient depletion that would arise from overuse, and to prevent the establishment of shrubs in the burn plots (which would decrease yields and reduce species diversity; Huffman et al. Reference Huffman, Stoddard, Springer, Crouse and Chancellor2013), each burn plot would have been used only once every three years, which is a rotational pattern that aligns with pinyon-juniper ecological succession patterns (Wagner et al. Reference Wagner, Smart, Ford, Trigg, Nichols and Smiley1984:617) and ethnographic accounts (Myers and Doolittle Reference Myers, Doolittle, Colton and Buckley2014:13; Roos Reference Roos, Mills and Fowles2017:689–690).

Figure 5. Randomly placed, spatially scaled anthropogenic niches (1.1 km in diameter), with insert showing embedded burn plots (400 m in diameter) by time period in the Upper Basin.

As a measure of the sustainability and low ecological impact of the fire foodway model, we predict that through time, burn plots within anthropogenic niches were deactivated and anthropogenic niches were abandoned altogether as population declined (Figure 6). For instance, if we focus on the yields for Amaranth 2 (Table 2), nine burn plots would have been established initially in nine anthropogenic niches between AD 875 and 1000. Thereafter, as population increased, the original nine anthropogenic niches would have been supplemented by 33 new ones during AD 1000–1050 and with 30 new ones during AD 1050–1075. After AD 1075, our figures indicate that no new anthropogenic niches needed to be established and, importantly, burning in previously established niches and their associated burn plots would have declined at an accelerating rate, thereby conserving the woodlands and their economic resources (cf. Peeples et al. Reference Peeples, Michael Barton and Schmich2006).

Figure 6. Dynamics of anthropogenic niche and burn plot establishment and abandonment in the Upper Basin through time.

The Archaeological Significance of the 2016 Scott Fire

Sometime shortly after midnight on June 28, 2016, a lightning strike ignited the Scott Fire in the western reaches of the Upper Basin (Figure 1). By the time the fire was suppressed on July 18, 2016, a total of 1,076.5 ha had burned, much of it so severely that the woodland was destroyed (Figure 7), a consequence of high fuel loads attributable to fire exclusion and suppression, among other factors (Fulé Reference Fulé, Stoffel, Bollschweiler, Butler and Luckman2010:376; Vankat Reference Vankat2013:278–281).

Figure 7. Upper Basin pinyon-juniper woodland (a) before (2008) and (b) after (2017) the lightning-caused Scott Fire (2016). (Color online)

The Scott Fire is significant for the fire foodway model in several respects. First, to provide a sense of scale in considering the effects of food fires on landscapes, the area burned by the Scott Fire is larger than the highest estimated number of hectares burned based on Amaranth 2 yields, 909 ha (Table 2). Second, as Figure 8 shows, by spring 2017, significant portions of the burned area were covered by “fetid goosefoot” (Dysphania graveolens), which historically has been an economically important plant to Pueblo peoples (Springer et al. Reference Springer, Daniels and Nazaire2009:324–325). The fact that this ruderal was particularly dense in and around abandoned masonry structures suggests not only that it is fire-responsive but that its seedbed has endured for centuries (Yarnell Reference Yarnell1965). Third, that the effects of the Scott Fire on the appearance of ruderals are not anomalous can be appreciated by considering the data in Table 3, which show a uniform ruderal response despite differences in fire size, ignition type, or severity.

Figure 8. Aftermath of the Scott Fire, which burned 1,076.5 ha in the Upper Basin between June 28 and July 18, 2016, showing dead trees and prehistoric masonry structure surrounded by fetid goosefoot in April 2017 (image used with permission of Neil Weintraub, Kaibab National Forest, US Department of Agriculture Forest Service).

Table 3. Forest Fires in the Upland American Southwest that Produced Amaranth or Chenopodium.

aOne to two years post-burn.

In addition to providing evidence for Yarnell's (Reference Yarnell1965) “Camp Follower” hypothesis, which is intended to explain abnormally high densities of wild economic plants in and around archaeological sites, the results of our study lend support for Henry Dobyns's (Reference Dobyns1972) “Altitude Sorting” hypothesis. Based on ethnohistoric accounts that describe the ubiquity of amaranth and chenopodium production in New World economies, widespread archaeological occurrences of cheno-ams in the American Southwest, and the environmental constraints on maize farming, particularly for densely occupied upland locales after AD 1000, Dobyns proposed that upland groups “depended upon amaranth and perhaps chenopodium cultivation,” in contrast to populations living in lower elevations that “grow more corns, beans, squash, and cotton, and less chenopodium and amaranth” (Reference Dobyns1972:45; see Bohrer Reference Bohrer1991:232–233 on the use of fire for ruderal production in the Sonoran Desert). The results of our and other modern archaeological investigations broadly align with the Camp Follower and Altitude Sorting hypotheses (e.g., Merrill et al. Reference Merrill, Hard, Mabry, Fritz, Adams, Roney and MacWilliams2009) and indicate that maize dependency was localized and uneven across the prehispanic American Southwest (Bayman and Sullivan Reference Bayman and Sullivan2008; Rocek Reference Rocek1995; cf. Spielmann et al. Reference Spielmann, Nelson, Ingram and Peeples2011).

Fire-Based Ruderal Production and the Economic Prehistory of the American Southwest

From southwestern Arizona (Bayman et al. Reference Bayman, Palacios-Fest, Fish and Huckell2004:132) to northeastern New Mexico (Kirkpatrick and Ford Reference Kirkpatrick and Ford1977) and from southeastern Nevada (McGuire et al. Reference McGuire, Hildebrandt, Gilreath, King and Berg2014) to southeastern New Mexico (Jelinek Reference Jelinek1966), archaeological evidence indicates that chenopodium and amaranth have been economic keystone species for centuries in the Southwest (Fritz et al. Reference Fritz, Bruno, Langlie, Smith, Kistler, Sayre and Bruno2017) and elsewhere in the New World (Carmody et al. Reference Carmody, Sherwood and Hoagland2017). The method of their cultivation, however, is poorly understood (Ford Reference Ford1981:22) because considerations of the economic effects of burning have not been coupled with the understanding that these plants are fire-responsive and that fire was an essential food-producing technology of prehistoric Southwestern societies. The plausibility of the fire-foodway model can be appreciated, we suggest, by embracing the idea that the effects of low-intensity burning on understory vegetation align with forest ecology, fire ecology, and the contents of the archaeobotanical record in ways that maize agriculture does not. Moreover, with human-controlled landscape fire, people actively manage the structure and composition of vegetation communities by disrupting ecological succession patterns, manipulating understory fuel loads, and rotating burn plots to encourage species diversity without soil nutrient loss (Bird Reference Bird2015). In short, the evidence indicates that sustainable fire-based ruderal agriculture was practiced extensively and in areas where maize agriculture was a risky and insecure endeavor (Dean Reference Dean, Tainter and Tainter1996).

We realize that some aspects of this model are ideal, if not mechanistic, constructs, for example, the size of burn plots, number of annual fires, and duration of fallow periods, but they serve to illustrate that, even when considering the lowest yields of just one plant and the caloric needs of the highest levels of population, fire foodways could easily provision the occupants of pinyon-juniper woodlands in areas that are environmentally hostile to maize farming. In addition, in view of the low impact of ruderal cultivation on pinyon-juniper woodlands, food fires could be ignited virtually continuously wherever sufficient fuels accumulate without jeopardizing the integrity of the ecosystem and its other economic resources, such as pinyon nuts and cactus (cf. Innes et al. Reference Innes, Blackford and Rowley-Conwy2013:88). In this regard, Lanner's observation that “the food potential of piñon forests in the Southwest has never been reliably estimated, but it is enormous” (Reference Lanner1981:105) is supported by our rough estimates of fire-responsive ruderal yields. Importantly, significant quantities of economic resources can be produced without much labor, without destroying the woodlands themselves, and by increasing edible biomass enormously, particularly if more than one ruderal was cultivated (Bye Reference Bye1981; Ford Reference Ford, Plog and Powell1984:135–137).

By incorporating aspects of the forest ecology and fire ecology of the Southwest's pinyon-juniper woodlands, our study illustrates the synergistic potential of applied historical ecology and niche construction theory in revealing the effects of anthropogenic fire in transforming the carrying capacity of these once heavily occupied ecosystems. Hence, it seems unlikely that the livelihoods of the prehistoric occupants of the Southwest's pinyon-juniper woodlands were challenged by slowly regenerating resources, by low carrying capacity, by scarce wild resources, and by the uncertain effectiveness of slash-and-burn maize cultivation (Wilcox Reference Wilcox and Ward1978). It is more ecologically realistic to view these woodlands as having been structured by the historic dynamics of anthropogenic fire regimes—systematic understory burning during periods of occupation, dramatic reduction of understory burning during periods of abandonment, and exclusion or suppression of fire during the twentieth century (Margolis Reference Margolis2014).Footnote 3

Conclusion

In archaeological contexts across the prehistoric American Southwest, the widespread ubiquity and high frequencies of ruderal plant remains—especially seeds and pollen from amaranth and chenopodium—support the proposition that the cultivation practices for these plants involved deliberately set understory fires (Sullivan and Forste Reference Sullivan and Forste2014). Furthermore, we think that anthropogenic fires, which triggered the disturbances that enabled ruderal production whenever and wherever sufficient understory fuels had accumulated, represent a form of sustainable agriculture whose yields were independent of the problematic rainfall and soil conditions that bedevil maize productivity and made it such a risky venture (Bocinsky and Varien Reference Bocinsky and Varien2017). The systematic management of understory fuel loads with low-intensity burning not only provided a dependable food supply, whose productive capacity was controlled by the ecosystem's inhabitants, but insulated the forests from catastrophic crown fires by eliminating the ladder-fuel problem that makes our overgrown woodlands so vulnerable to wildfires today (Huffman et al. Reference Huffman, Stoddard, Springer, Crouse and Chancellor2013), as the Scott Fire dramatically demonstrates.

Ruderal-producing fire foodways were sustainable because they required minimal fuel loads and efforts to ignite them, could be prosecuted on the most agriculturally unproductive soils, and, in the case of amaranth specifically, did well with less than 25 cm of precipitation and did not deplete soil of nitrogen to the same degree as maize. Intercropped with other drought-resistant, fire-responsive ruderals, such as chenopodium and various grasses, low-intensity understory “food fires” created a foodway that was largely inoculated against long-term climatic change and short-term environmental variability (Fulé et al. Reference Fulé, Wallace Covington, Moore, Heinlein and Waltz2002a:44). Moreover, fire foodways take advantage of highly predictable successional pathways (Barney and Frischknecht Reference Barney and Frischknecht1974) that virtually guaranteed, with fire rotation (i.e., resting formerly burned areas for a year or two to promote fuel regeneration; Lightfoot and Cuthrell Reference Lightfoot and Cuthrell2015:1585), interannual continuity in food provisioning (Bates and Davies Reference Bates and Davies2016:127). Hence, this sustainable strategy contributed to a degree of food security that was rarely enjoyed by maize-based foodways in view of their vulnerability to capricious and largely uncontrollable environmental factors. The fire-foodway model illustrates that, by integrating archaeobotanical, fire-ecological, and surface archaeological data, we can enrich our narratives about how groups of economically autonomous people, with the judicious application of fire, engineered landscapes and secured livelihoods by unleashing the productivity capacity of ancient pinyon-juniper woodlands—upland ecosystems whose fire regimes, structure, and appearance have few counterparts in the twenty-first century.

Acknowledgments

Financial support for the Upper Basin Archaeological Research Project has been provided by the US Department of Agriculture Forest Service, the National Center for Preservation Technology and Training (National Park Service), the National Geographic Society/Waitt Grants Program, the C. P. Taft Research Center (University of Cincinnati), and the University Research Council (University of Cincinnati). We are grateful to Dr. Maria Nieves Zedeño (University of Arizona) and to Dr. Jeff Jacobson (University of Cincinnati) for translating our abstract into Spanish. Archaeological investigations in Grand Canyon National Park during 2012 and 2014 were authorized by permits GRCA-2012-SCI-0006 and GRCA-2014-SCI-0033. Federal housing occupancy and archaeological investigations in Kaibab National Forest were authorized by Special Use Permits TUS85-352 (2012) and KAI105 (2014), respectively.

Data Availability Statement

Specialist reports of macrobotanical and pollen analyses, data files of archaeological survey records (including differentially corrected GPS data), and GIS shapefiles of Upper Basin environmental and archaeological characteristics used in the preparation of this article are available by contacting the corresponding author. In addition, primary archaeobotanical data referred to in this essay, including sample collection and analysis procedures, context types, sample numbers, identified taxa, nomenclature protocols, plant part, specimen condition, the presence or absence of charring, and miscellaneous comments, may be found by consulting the following sources: Berkebile (Reference Berkebile2014); Brandt (Reference Brandt1991); Cummings and Puseman (Reference Cummings and Puseman1995, Reference Cummings and Puseman1997, Reference Cummings, Puseman and Downum2010); Huckell (Reference Huckell and Whittlesey1992); Scott (Reference Scott and Sullivan1986).

Supplemental Materials

For supplemental material accompanying this article, visit https://doi.org/10.1017/aaq.2018.32.

Supplemental Table 1. Archaeological Characteristics of 10 Excavated Sites in Grand Canyon National Park and the Upper Basin, Northern Arizona, that Produced Macrobotanical Samples.

Supplemental Table 2. Productivity Estimates of Amaranth and Chenopodium.

Supplemental Text 1. Key References Consulted about the Relation between Anthropogenic Fire and Human Prehistory.

Supplemental Text 2. Key References Consulted about the Relation between Anthropogenic Landscape Fire Ecology and Subsistence Economies.

Supplemental Text 3. Key Ethnographic, Ethnohistoric, and Cross-Cultural Studies Consulted about Anthropogenic Landscape Burning.

Supplemental Text 4. Key References Consulted about Maize and Amaranth Evolution and Cultivation in the New World.

Supplemental Text 5. Population and Amaranth Productivity Estimates for the Upper Basin, AD 875–1200.

Footnotes

1. Estimates range from 47,133 km2 in Arizona alone (Miller and Tausch Reference Miller, Tausch, Galley and Wilson2001:16) to 153,100 km2 (Vankat Reference Vankat2013:268) to 291,374 km2 (West Reference West, Anderson, Fralish and Baskin1999:288) for the greater Southwest.

2. Given the extensive interannual variability in paleoprecipitation patterns, there is no correlation between the precipitation values for a given year and its prior year (r = 0.025, p = 0.663, n = 301). Also, the results of a nonparametric Runs test (number of runs = 164) indicate that annual precipitation values (n = 301) are randomly distributed using either the mean ( = 36.6 cm, Z = 1.46, p = 0.144) or median (Md = 36.2 cm, Z = 1.44, p = 0.149; Thomas Reference Thomas1986:339–340). (Data provided courtesy of Jeffrey S. Dean, Laboratory of Tree-Ring Research, University of Arizona, Tucson, personal communication 2001.)

3. The results of grain size, soil phosphorous (P), charcoal concentration, and carbon isotope ratio analyses of axial alluvial fan deposits (McNamee Reference McNamee2003; Roos et al. Reference Roos, Sullivan, McNamee and Dean2010) and from botanical and pollen analyses of samples recovered from a wide variety of archaeological contexts (Sullivan and Ruter Reference Sullivan, Ruter, Doyel and Dean2006; Sullivan et al. Reference Sullivan, Berkebile, Forste and Washam2015) support the argument that the prehistoric fire regime in the Upper Basin between circa AD 875 and 1200 consisted of frequent, low-intensity, surface, understory anthropogenic fires (West Reference West1984:1310). Mixed-severity or high-severity (lethal) fires (where 20% to more than 70% of the overstory is killed by fire [Williams and Baker Reference Williams and Baker2013:301]) were rare until the area was abandoned (i.e., perennial occupation ceased ca. AD 1200). In contrast, the fire regime between AD 1700 and 2000, based on fire scar records (Fulé et al. Reference Fulé, Heinlein, Wallace Covington and Moore2003) and analysis of historic forest structure (Fulé et al. Reference Fulé, Wallace Covington, Smith, Springer, Heinlein, Huisinga and Moore2002b; Williams and Baker Reference Williams and Baker2013), is characterized by infrequent, low- to medium-intensity, surface, nonanthropogenic fires punctuated by periodic severe (stand-replacing) canopy fires (Fulé Reference Fulé, Stoffel, Bollschweiler, Butler and Luckman2010; cf. Liebmann et al. Reference Liebmann, Farella, Roos, Stack, Martini and Swetnam2016). The character of the fire regime between AD 1200 and 1700 is unknown save for the presumption that fuel loads and ignitions were unmanaged by humans (cf. Herring et al. Reference Herring, Scott Anderson and San Miguel2014:860).

References

References Cited

Abrams, Marc D., and Nowacki, Gregory J. 2008 Native Americans as Active and Passive Promoters of Mast and Fruit Trees in the Eastern USA. Holocene 18:11231137.Google Scholar
Adams, Karen R. 2011 Anthropogenic Ecology in the American Southwest. In Movement, Connectivity, and Landscape Change in the Ancient Southwest, edited by Nelson, Margaret C. and Strawhacker, Colleen, pp. 119140. University Press of Colorado, Boulder.Google Scholar
Adams, Karen R. 2015 The Archaeology and Agronomy of Ancient Maize (Zea mays L.). In Traditional Arid Lands Agriculture: Understanding the Past for the Future, edited by Ingram, Scott E. and Hunt, Robert C., pp. 1553. University of Arizona Press, Tucson.Google Scholar
Adams, Karen R., and Dockter, Abigail R. 2013 Five Years of Fire Recovery in Mesa Verde National Park. Manuscript on file, Crow Canyon Archaeological Center, Cortez, Colorado.Google Scholar
Adams, Karen R., and Fish, Suzanne K. 2011 Subsistence through Time in the Greater Southwest. In The Subsistence Economies of Indigenous North American Societies: A Handbook, edited by Smith, Bruce D., pp. 147183. Smithsonian Institution Scholarly Press, Washington, DC.Google Scholar
Agogino, George 1960 The San Jose Sites: A Cochise-Like Manifestation in the Middle Rio Grande. Southwestern Lore 26:4348.Google Scholar
Allen, Craig D. 2002 Lots of Lightning and Plenty of People: An Ecological History of Fire in the Upland Southwest. In Fire, Native People, and the Natural Landscape, edited by Vale, Thomas R., pp. 143193. Island Press, Washington, DC.Google Scholar
Anscheutz, Kurt F. 2006 Tewa Fields, Tewa Traditions. In Canyon Gardens: The Ancient Pueblo Landscapes of the American Southwest, edited by Price, V. B. and Morrow, Baker H., pp. 5773. University of New Mexico Press, Albuquerque.Google Scholar
Barney, Milo A., and Frischknecht, Neil C. 1974 Vegetation Changes Following Fire in the Pinyon-Juniper Type of West-Central Utah. Journal of Range Management 27:9196.Google Scholar
Bates, Jonathan D., and Davies, Kirk W. 2016 Seasonal Burning of Juniper Woodlands and Spatial Recovery of Herbaceous Vegetation. Forest Ecology and Management 361:117130.Google Scholar
Bayman, James M., Palacios-Fest, Manuel R., Fish, Suzanne K., and Huckell, Lisa W. 2004 The Paleoecology and Archaeology of Long-Term Water Storage in a Hohokam Reservoir, Southwestern Arizona, U.S.A. Geoarchaeology: An International Journal 19:119140.Google Scholar
Bayman, James M., and Sullivan, Alan P. III 2008 Property, Identity, and Macroeconomy in the Prehispanic Southwest. American Anthropologist 110:115.Google Scholar
Begay, Richard M., and Roberts, Alexandra 1996 The Early Navajo Occupation of the Grand Canyon Region. In The Archaeology of Navajo Origins, edited by Towner, Ronald H., pp. 197210. University of Utah Press, Salt Lake City.Google Scholar
Bellorado, Benjamin A., and Anderson, Kirk C. 2013 Early Pueblo Responses to Climate Variability: Farming Traditions, Land Tenure, and Social Power in the Eastern Mesa Verde Region. Kiva 78:377416.Google Scholar
Benson, Larry V. 2011 Factors Controlling Pre-Columbian and Early Historic Maize Productivity in the American Southwest, Part I: The Southern Colorado Plateau and Rio Grande Regions. Journal of Archaeological Method and Theory 18:160.Google Scholar
Benson, Larry V., Ramsey, D. K., Stahle, D. W., and Petersen, K. L. 2013 Some Thoughts on the Factors that Controlled Prehistoric Maize Production in the American Southwest with Application to Southwestern Colorado. Journal of Archaeological Science 40:28692880.Google Scholar
Berkebile, Jean N. 2014 Investigating Subsistence Diversity in the Upper Basin: A Second Look at Archaeobotanical Remains from MU 125, a Late Pueblo II Settlement. Master's thesis, Department of Anthropology, University of Cincinnati, Cincinnati, Ohio.Google Scholar
Berlin, G. Lennis, Salas, David E., and Geib, Phil R. 1990 A Prehistoric Sinagua Agricultural Site in the Ashfall Zone of Sunset Crater, Arizona. Journal of Field Archaeology 17:116.Google Scholar
Bird, Rebecca B. 2015 Disturbance, Complexity, Scale: New Approaches to the Study of Human-Environment Interactions. Annual Review of Anthropology 44:241257.Google Scholar
Bocinsky, R. Kyle, and Varien, Mark D. 2017 Comparing Maize Paleoproduction Models with Experimental Data. Journal of Ethnobiology 37:282307.Google Scholar
Bohrer, Vorsila L. 1962 Nature and Interpretation of Ethnobotanical Materials from Tonto National Monument. In Archeological Studies at Tonto National Monument, Arizona, edited by Caywood, Louis R., pp. 75114. Southwestern Monuments Association, Globe, Arizona.Google Scholar
Bohrer, Vorsila L. 1972 Paleoecology of the Hay Hollow Site, Arizona. Fieldiana Anthropology 63:130.Google Scholar
Bohrer, Vorsila L. 1975 The Prehistoric and Historic Role of the Cool-Season Grasses in the Southwest. Economic Botany 29:199207.Google Scholar
Bohrer, Vorsila L. 1991 Recently Recognized Cultivated and Encouraged Plants among the Hohokam. Kiva 56:227235.Google Scholar
Bohrer, Vorsila L. 2007 Preceramic Subsistence in Two Rock Shelters in Fresnal Canyon, South Central New Mexico. Arizona State Museum Archaeological Series 199. University of Arizona, Tucson.Google Scholar
Bond, William J., and Keeley, Jon E. 2005 Fire as a Global “Herbivore:” The Ecology and Evolution of Flammable Ecosystems. Trends in Ecology and Evolution 20:388394.Google Scholar
Bonnicksen, Thomas M. 2000 America's Ancient Forests: From the Ice Age to the Age of Discovery. John Wiley and Sons, New York.Google Scholar
Bowman, David M. J. S., Balch, Jennifer K., Artaxo, Paulo, Bond, William J., Carlson, Jean M., Cochrane, Mark A., D'Antonio, Carla M., DeFries, Ruth S., Doyle, John C., Harrison, Sandy P., Johnston, Fay H., Keeley, Jon E., Krawchuk, Meg A., Kull, Christian A., Brad Marston, J., Moritz, Max A., Colin Prentice, I., Roos, Christopher I., Scott, Andrew C., Swetnam, Thomas W., van der Werf, Guido R., and Pyne, Stephen J. 2009 Fire in the Earth System. Science 234:481484.Google Scholar
Bozarth, Steven 1992 Fossil Pollen and Phytolith Analysis. In Archaeological Investigations at Lee Canyon: Kayenta Anasazi Farmsteads in the Upper Basin, Coconino County, Arizona, edited by Whittlesey, Stephanie M., pp. 135144. Technical Series No. 38. Statistical Research, Tucson, Arizona.Google Scholar
Braje, Todd J. 2015 Earth Systems, Human Agency, and the Anthropocene: Planet Earth in the Human Age. Journal of Archaeological Research 23:369396.Google Scholar
Brandt, Carol B. 1991 Report of Flotation Analysis for Sites MU 38, MU 235, and MU 236. Prepared for the Department of Anthropology, University of Cincinnati, Cincinnati, Ohio. Copies available from Zuni Archaeological Program, Zuni, New Mexico.Google Scholar
Brewer, David G., Jorgensen, Rodney K., Munk, Lewis P., Robbie, Wayne A., and Travis, Janet L. 1991 Terrestrial Ecosystems Survey of the Kaibab National Forest. US Department of Agriculture Forest Service, Albuquerque, New Mexico.Google Scholar
Bye, Robert A. Jr. 1981 Quelites—Ethnoecology of Edible Greens—Past, Present, and Future. Journal of Ethnobiology 1:109123.Google Scholar
Carmody, Stephen B., Sherwood, Sarah C., and Hoagland, Carolyn 2017 From the Past … a More Sustainable Future? SAA Archaeological Record 17 (March): 1016.Google Scholar
Cleeland, Terri, Hanson, John, Lesko, Lawrence, and Weintraub, Neil 1990 Ethnographic Use of the South Kaibab. Manuscript on file, Kaibab National Forest, Williams, Arizona.Google Scholar
Crabtree, Stefani A., Vaughn, Lydia J. S., and Crabtree, Nathan T. 2017 Reconstructing Ancestral Pueblo Food Webs in the Southwestern United States. Journal of Archaeological Science 81:116127.Google Scholar
Cummings, Linda Scott 1995 Agriculture and the Mesa Verde Area Anasazi Diet: Description and Nutritional Analysis. In Soil, Water, Biology, and Belief in Prehistoric and Traditional Southwestern Agriculture, edited by Wolcott Toll, H., pp. 335352. Special Publication No. 2. New Mexico Archaeological Council, Albuquerque.Google Scholar
Cummings, Linda Scott, and Puseman, Kathryn 1995 Pollen and Macrofloral Analysis at Sites MU 123, MU 125, and MU 235 in Kaibab National Forest, North-Central Arizona. Technical Report 95-10. Prepared for the Department of Anthropology, University of Cincinnati, Cincinnati, Ohio. Copies available from Paleo Research Laboratories, Denver, Colorado.Google Scholar
Cummings, Linda Scott, and Puseman, Kathryn 1997 Pollen and Macrofloral Analysis at Sites MU 125 and MU 125A, Arizona. Technical Report 97-45. Prepared for the Department of Anthropology, University of Cincinnati, Cincinnati, Ohio. Copies available from Paleo Research Laboratories, Denver, Colorado.Google Scholar
Cummings, Linda Scott, and Puseman, Kathryn 2010 Pollen, Macrofloral, and Organic Residue (FTIR) Analysis for Site AZ B:16:105, Northern Arizona. In Archaeological Excavations at Site B:16:105, Grand Canyon National Park, by Downum, Christian E., pp. 83115. Northern Arizona University Archaeological Report 1247. Flagstaff.Google Scholar
Darling, Marilyn L. S. 1967 Structure and Productivity of Pinyon-Juniper Woodland in Northern Arizona. PhD dissertation, Department of Zoology, Duke University, Durham, North Carolina.Google Scholar
Davis, Owen K. 1986 Pollen Analysis from AZ:I:1:17 (ASM). In Prehistory of the Upper Basin, Coconino County, Arizona, edited by Sullivan, Alan P. III, pp. 333338. Arizona State Museum Archaeological Series No. 167. University of Arizona, Tucson.Google Scholar
Dean, Jeffrey S. 1996 Demography, Environment, and Subsistence Stress. In Evolving Complexity and Environmental Risk in the Prehistoric Southwest, edited by Tainter, Joseph A. and Tainter, Bonnie B., pp. 2558. Santa Fe Institute Studies in the Sciences of Complexity Vol. XXIV. Addison-Wesley, Reading, Massachusetts.Google Scholar
Dobyns, Henry F. 1972 Altitude Sorting of Ethnic Groups in the Southwest. Plateau 47:4248.Google Scholar
Dods, Roberta R. 2002 The Death of Smokey Bear: The Ecodisaster Myth and Forest Management Practices in Prehistoric North America. World Archaeology 33:475487.Google Scholar
Doebley, John F. 1981 Plant Remains Recovered by Flotation from Trash at Salmon Ruin, New Mexico. Kiva 46:168187.Google Scholar
Doolittle, William E. 2000 Cultivated Landscapes of Native North America. Oxford University Press, New York.Google Scholar
Downum, Christian E., and Sullivan, Alan P. III 1990 Description and Analysis of Prehistoric Settlement Patterns at Wupatki National Monument. In The Wupatki Archaeological Inventory Survey Project, compiled by Anderson, Bruce A., pp. 5.15.90. Southwest Cultural Resources Center Professional Paper No. 35. National Park Service, Santa Fe, New Mexico.Google Scholar
Effland, Richard W. Jr., Jones, Anne T., and Euler, Robert C. 1981 The Archaeology of Powell Plateau: Regional Interaction at Grand Canyon. Grand Canyon Natural History Association, Grand Canyon, Arizona.Google Scholar
Erdman, James A. 1970 Pinyon-Juniper Succession after Natural Fires on Residual Soils of Mesa Verde, Colorado. Biological Series No. 9. Brigham Young University, Provo, Utah.Google Scholar
Euler, Robert C. 1988 Demography and Cultural Dynamics on the Colorado Plateau. In The Anasazi in a Changing Environment, edited by Gumerman, George J., pp. 192229. Cambridge University Press, Cambridge.Google Scholar
Euler, Robert C. 1992 Grand Canyon Indians. In The Grand Canyon: Intimate Views, edited by Euler, Robert C. and Tikalsky, Frank, pp. 4360. University of Arizona Press, Tucson.Google Scholar
Everett, Richard L., and Ward, Kenneth 1984 Early Plant Succession on Pinyon-Juniper Controlled Burns. Northwest Science 58:5768.Google Scholar
Floyd, M. Lisa, Romme, William H., and Hanna, David D. 2003 Fire History. In Ancient Pinyon-Juniper Woodlands: A Natural History of Mesa Verde Country, edited by Lisa Floyd, M., pp. 261277. University Press of Colorado, Boulder.Google Scholar
Ford, Richard I. 1981 Gardening and Farming before A.D. 1000: Patterns of Prehistoric Cultivation North of Mexico. Journal of Ethnobiology 1:627.Google Scholar
Ford, Richard I. 1984 Ecological Consequences of Early Agriculture in the Southwest. In Papers on the Archaeology of Black Mesa, Arizona, Vol. II, edited by Plog, Stephen and Powell, Shirley, pp. 127138. Southern Illinois University Press, Carbondale.Google Scholar
Ford, Richard I. 2000 Human Disturbance and Biodiversity: A Case Study from Northern New Mexico. In Biodiversity and Native America, edited by Minnis, Paul E. and Elisens, Wayne J., pp. 207222. University of Oklahoma Press, Norman.Google Scholar
Foust, Nathaniel E. 2015 A Spatiotemporal GIS Analysis of GPS Effects on Archaeological Site Variability. Master's thesis, Department of Anthropology, University of Cincinnati, Cincinnati, Ohio.Google Scholar
Foxx, Teralene S. 1996 Vegetation Succession after the La Mesa Fire at Bandelier National Monument. In Fire Effects in Southwestern Forests: Proceedings of the Second La Mesa Fire Symposium, edited by Allen, Craig D., pp. 4769. US Department of Agriculture Forest Service, Fort Collins, Colorado.Google Scholar
French, Charles, Perriman, Richard, Cummings, Linda S., Hall, Stephen, Goodman-Elgar, Melissa, and Boreham, Julie 2009 Holocene Alluvial Sequences, Cumulic Soils, and Fire Signatures in the Middle Rio Puerco Basin at Guadalupe Ruin, New Mexico. Geoarchaeology: An International Journal 24:638676.Google Scholar
Fritz, Gayle J. 2007 Pigweeds for the Ancestors: Cultural Identities and Archaeological Identification Methods. In The Archaeology of Food and Identity, edited by Twiss, Katheryn C., pp. 288307. Center for Archaeological Investigations Occasional Paper No. 34. Southern Illinois University, Carbondale.Google Scholar
Fritz, Gayle J., Adams, Karen R., Rice, Glen E., and Czarzasty, John L. 2009 Evidence for Domesticated Amaranth (Amaranthus) from a Sedentary Period Hohokam House Floor at Las Canopas. Kiva 75:393418.Google Scholar
Fritz, Gayle J., Bruno, Maria C., Langlie, Brieanna S., Smith, Bruce D., and Kistler, Logan 2017 Cultigen Chenopods in the Americas: A Hemispherical Perspective. In Social Perspectives on Ancient Lives from Paleoethnobotanical Data, edited by Sayre, Matthew P. and Bruno, Maria C., pp. 5575. Springer International, Dordrecht, the Netherlands.Google Scholar
Fulé, Peter Z. 2010 Wildfire Ecology and Management at Grand Canyon, AZ: Tree-Ring Applications in Forest Fire History and Modeling. In Tree Rings and Natural Hazards: A State-of-the-Art, edited by Stoffel, Markus, Bollschweiler, Michelle, Butler, David R., and Luckman, Brian H., pp. 365381. Springer, New York.Google Scholar
Fulé, Peter Z., Wallace Covington, W., Moore, Margaret M., Heinlein, Thomas A., and Waltz, Amy E. M. 2002a Natural Variability in Forests of the Grand Canyon, USA. Journal of Biogeography 29:3147.Google Scholar
Fulé, Peter Z., Wallace Covington, W., Smith, H. B., Springer, Judith D., Heinlein, Thomas A., Huisinga, Kristin D., and Moore, Margaret M. 2002b Comparing Ecological Restoration Alternatives: Grand Canyon, Arizona. Forest Ecology and Management 170:1941.Google Scholar
Fulé, Peter Z., Heinlein, Thomas A., Wallace Covington, W., and Moore, Margaret M. 2003 Assessing Fire Regimes on Grand Canyon Landscapes with Fire-Scar and Fire-Record Data. International Journal of Wildland Fire 12:129145.Google Scholar
Han, Y. M., Peteet, D. M., Arimoto, R., Cao, J. J., An, Z. S., Sritrairat, S., and Yan, B. Z. 2016 Climate and Fuel Controls on North American Paleofires: Smoldering to Flaming in the Late-Glacial-Holocene Transition. Nature Scientific Reports 6:20719. https://www.nature.com/articles/srep20719, accessed September 22, 2016.Google Scholar
Haws, Jonathan A. 2012 Paleolithic Socionatural Relationships during MIS 3 and 2 in Central Portugal. Quaternary International 264:6177.Google Scholar
Hendricks, David M. 1985 Arizona Soils. College of Agriculture, University of Arizona, Tucson.Google Scholar
Herring, Erin M., Scott Anderson, R., and San Miguel, George L. 2014 Fire, Vegetation, and Ancestral Puebloans: A Sediment Record from Prater Canyon in Mesa Verde National Park, Colorado, USA. Holocene 24:853863.Google Scholar
Hill, James N., and Hevly, Richard H. 1968 Pollen at Broken K Pueblo: Some New Interpretations. American Antiquity 33:200210.Google Scholar
Homburg, Jeffrey A. 1992 Soil Fertility Study. In Archaeological Investigations at Lee Canyon: Kayenta Anasazi Farmsteads in the Upper Basin, Coconino County, Arizona, edited by Whittlesey, Stephanie M., pp. 145161. Technical Series No. 38. Statistical Research, Tucson, Arizona.Google Scholar
Hough, Ian, and Brennan, Ellen 2008 Architectural Documentation and Preservation of Havasupai and Navajo Wooden Pole Structures in Grand Canyon National Park. In Reflections of Grand Canyon Historians: Ideas, Arguments, and First-Person Accounts, edited by Berger, Todd R., pp. 8188. Grand Canyon Association, Grand Canyon, Arizona.Google Scholar
Huckell, Lisa W. 1992 Plant Remains. In Archaeological Investigations at Lee Canyon: Kayenta Anasazi Farmsteads in the Upper Basin, Coconino County, Arizona, edited by Whittlesey, Stephanie M., pp. 119133. Technical Series No. 38. Statistical Research, Tucson, Arizona.Google Scholar
Huffman, D. W., Stoddard, M. T., Springer, J. D., Crouse, J. E., and Chancellor, W. W. 2013 Understory Plant Community Responses to Hazardous Fuels Reduction Treatments in Pinyon-Juniper Woodlands of Arizona, USA. Forest Ecology and Management 289:478488.Google Scholar
Huffman, Mary R. 2013 The Many Elements of Traditional Fire Knowledge: Synthesis, Classification, and Aids to Cross-Cultural Problem Solving in Fire-Dependent Systems around the World. Ecology and Society 18(4). DOI:10.5751/ES-05843-180403, accessed October 15, 2016.Google Scholar
Huisinga, Kristin D., Laughlin, Daniel L., Fulé, Peter Z., Springer, Judith D., and McGlone, Christopher M. 2005 Effects of an Intense Prescribed Fire on Understory Vegetation in a Mixed Conifer Forest. Journal of the Torrey Botanical Society 134:590601.Google Scholar
Hunt, C. O., and Rabett, R. J. 2014 Holocene Landscape Intervention and Plant Food Production Strategies in Island and Mainland Southeast Asia. Journal of Archaeological Science 51:2233.Google Scholar
Hunter, Andrea A. 1997 Seeds, Cucurbits, and Corn from Lizard Man Village. Kiva 62:221244.Google Scholar
Innes, James B., Blackford, Jeffrey J., and Rowley-Conwy, Peter A. 2013 Late Mesolithic and Early Neolithic Forest Disturbance: A High Resolution Paleoecological Test of Human Impact Hypotheses. Quaternary Science Reviews 77:80100.Google Scholar
Jelinek, Arthur J. 1966 Correlation of Archaeological and Palynological Data. Science 152:15071509.Google Scholar
Jones, Volney H. 1953 Review of “The Grain Amaranths: A Survey of Their History and Classification” by Jonathan D. Sauer. American Antiquity 19:9092.Google Scholar
Jones, Volney H., and Fonner, Robert L. 1954 Plant Materials from Sites in the Durango and La Plata Areas, Colorado. In Basket Maker II Sites near Durango, Colorado, by Morris, Earl H. and Burgh, Robert F., pp. 93115. Publication 604. Carnegie Institution of Washington, Washington, DC.Google Scholar
Kirkpatrick, David T., and Ford, Richard I. 1977 Basketmaker Food Plants from the Cimarron District, Northeastern New Mexico. Kiva 42:257269.Google Scholar
Kohler, Timothy A. 2012 Modeling Agricultural Productivity and Farming Effort. In Emergence and Collapse of Early Villages: Models of Central Mesa Verde Archaeology, edited by Kohler, Timothy A. and Varien, Mark D., pp. 85111. University of California Press, Berkeley.Google Scholar
Kohler, Timothy A., Kyle Bocinsky, R., Cockburn, Denton, Crabtree, Stefani A., Varien, Mark D., Kolm, Kenneth E., Smith, Schaun, Ortman, Scott G., and Kobti, Ziab 2012 Modelling Prehispanic Pueblo Societies in Their Ecosystems. Ecological Modelling 41:3041.Google Scholar
Kohler, Timothy A., Gumerman, George J., and Reynolds, Robert G. 2005 Simulating Ancient Societies. Scientific American 293:7784.Google Scholar
Kohler, Timothy A., Kresl, James, West, Carla Van, Carr, Eric, and Wilshusen, Richard H. 2000 Be There Then: A Modeling Approach to Settlement Determinants and Spatial Efficiency among Late Ancestral Pueblo Populations of the Mesa Verde Region, U.S. Southwest. In Dynamics in Human and Primate Societies, edited by Kohler, Timothy A. and Gumerman, George J., pp. 145178. Oxford University Press, Oxford.Google Scholar
Kohler, Timothy A., and Van West, Carla R. 1996 The Calculus of Self-Interest in the Development of Cooperation: Sociopolitical Development and Risk among the Northern Anasazi. In Evolving Complexity and Environmental Risk in the Prehistoric Southwest, edited by Tainter, Joseph A. and Tainter, Bonnie B., pp. 169196. Santa Fe Institute Studies in the Sciences of Complexity Vol. XXIV. Addison-Wesley, Reading, Massachusetts.Google Scholar
Kuenzi, Amanda M., Fulé, Peter Z., and Sieg, Carolyn H. 2008 Effects of Fire Severity and Pre-Fire Stand Treatment on Plant Community Recovery after a Large Wildfire. Forest Ecology and Management 255:855865.Google Scholar
Lanner, Ronald M. 1981 The Piñon Pine: A Natural and Cultural History. University of Nevada Press, Reno.Google Scholar
Laughlin, Daniel C., Bakker, Jonathan D., Stoddard, Michael T., Daniels, Mark L., Springer, Judith D., Gildar, Cata N., Green, Aaron M., and Wallace Covington, W. 2004 Toward Reference Conditions: Wildfire Effects on Flora in an Old-Growth Ponderosa Pine Forest. Forest Ecology and Management 199:137152.Google Scholar
Liebmann, Matthew J., Farella, Joshua, Roos, Christopher I., Stack, Adam, Martini, Sarah, and Swetnam, Thomas W. 2016 Native American Depopulation, Reforestation, and Fire Regimes in the Southwest United States, 1492–1900 CE. Proceedings of the National Academy of Sciences 113:E696E704.Google Scholar
Lightfoot, Kent G., and Cuthrell, Rob Q. 2015 Anthropogenic Burning and the Anthropocene in Late-Holocene California. Holocene 25:15811587.Google Scholar
Lightfoot, Kent G., Cuthrell, Rob Q., Striplen, Chuck J., and Hylkema, Mark G. 2013 Rethinking the Study of Landscape Management Practices among Hunter-Gatherers in North America. American Antiquity 78:285301.Google Scholar
Lindsay, B. A., Strait, R. K., and Denny, D. W. 2003 Soil Survey of the Grand Canyon Area, Arizona, Parts of Coconino and Mohave Counties. US Department of Agriculture Natural Resources Conservation Service, Phoenix, Arizona.Google Scholar
McGuire, Kelly, Hildebrandt, William, Gilreath, Amy, King, Jerone, and Berg, John 2014 The Prehistory of Gold Butte: A Virgin River Hinterland, Clark County, Nevada. Anthropological Papers 127. University of Utah Press, Salt Lake City.Google Scholar
McNamee, Calla 2003 A Geoarchaeological Investigation of Simkins Flat, Upper Basin, Northern Arizona. Master's thesis, Department of Archaeology, University of Calgary, Alberta, Canada.Google Scholar
Margolis, Ellis Q. 2014 Fire Regime Shift Linked to Increased Forest Density in a Pinyon-Juniper Savanna Landscape. International Journal of Wildland Fire. DOI:10.1071/WF13053, accessed August 20, 2017.Google Scholar
Martin, John F. 1985 The Prehistory and Ethnohistory of Havasupai-Hualapai Relations. Ethnohistory 32:135153.Google Scholar
Mason, S. L. R. 2000 Fire and Mesolithic Subsistence—Managing Oaks for Acorns in Northwest Europe? Palaeogeography, Palaeoclimatology, Palaeoecology 164:139150.Google Scholar
Matson, R. G., Lipe, William D., and Haase, William R. IV 1988 Adaptational Continuities and Occupational Discontinuities: The Cedar Mesa Anasazi. Journal of Field Archaeology 15:245264.Google Scholar
Mellars, Paul 1976 Fire Ecology, Animal Populations and Man: A Study of Some Ecological Relationships in Prehistory. Proceedings of the Prehistoric Society 42:1545.Google Scholar
Merkle, John 1952 An Analysis of a Pinyon-Juniper Community at Grand Canyon, Arizona. Ecology 33:375384.Google Scholar
Merrill, William L., Hard, Robert J., Mabry, Jonathan B., Fritz, Gayle L., Adams, Karen R., Roney, John R., and MacWilliams, A. C. 2009 The Diffusion of Maize to the Southwestern United States and Its Impact. Proceedings of the National Academy of Sciences 106:2101921026.Google Scholar
Metzger, D. G. 1961 Geology in Relation to Availability of Water along the South Rim, Grand Canyon National Park, Arizona. US Government Printing Office, Washington, DC.Google Scholar
Miller, Richard F., and Tausch, Robin J. 2001 The Role of Fire in Juniper and Pinyon Woodlands: A Descriptive Analysis. In Proceedings of the Invasive Species Workshop: The Role of Fire in the Control and Spread of Invasive Species, edited by Galley, K. E. M. and Wilson, T. P., pp. 1530. Miscellaneous Publication No. 11. Tall Timbers Research Station, Tallahassee, Florida.Google Scholar
Mink, Philip B. II 2015 Living on the Edge: Rethinking Pueblo Period (AD 700–1225) Indigenous Settlement Patterns within Grand Canyon National Park, Northern Arizona. PhD dissertation, Department of Anthropology, University of Kentucky, Lexington.Google Scholar
Minnis, Paul E. 1991 Famine Foods of the Northern American Southwest Desert Borderlands in Historical Context. Journal of Ethnobiology 11:231257.Google Scholar
Morrow, Baker H. 2006 The Berry Gardens of Quarai and the Pocket Terraces of Abo. In Canyon Gardens: The Ancient Pueblo Landscapes of the American Southwest, edited by Price, V. B. and Morrow, Baker H., pp. 1731. University of New Mexico Press, Albuquerque.Google Scholar
Myers, Michael D., and Doolittle, William E. 2014 The New Narrative on Native Landscape Transformations. In North American Odyssey: Historical Geographies for the Twenty-First Century, edited by Colton, Craig E. and Buckley, Geoffrey L., pp. 926. Rowan and Littlefield, Lanham, Maryland.Google Scholar
Nabhan, Gary P., Coder, Marcelle, and Smith, Susan J. 2004 Woodlands in Crisis: A Legacy of Lost Diversity on the Colorado Plateau. Bilby Research Center Occasional Papers No. 7. Northern Arizona University, Flagstaff.Google Scholar
Ortman, Scott G., Diederichs, Shanna, Schleher, Kari, Fetterman, Jerry, Espinosa, Marcus, and Sommer, Caitlin 2016 Demographic and Social Dimensions of the Neolithic Revolution in Southwest Colorado. Kiva 82:232258.Google Scholar
Peeples, Matthew A., Michael Barton, C., and Schmich, Steven 2006 Resilience Lost: Intersecting Land Use and Landscape Dynamics in the Prehistoric Southwestern United States. Ecology and Society 11. Electronic document, https://www.ecologyandsociety.org/vol11/iss2/art22/, accessed December 10, 2011.Google Scholar
Plog, Stephen, Fish, Paul R., Glowacki, Donna M., and Fish, Suzanne K. 2015 Key Issues and Topics in the Archaeology of the American Southwest and Northwestern Mexico. Kiva 81:230.Google Scholar
Pool, Michael D. 2013 Mimbres Mogollon Farming: Estimating Prehistoric Agricultural Production during the Classic Mimbres Period. In Soils, Climate and Societies: Archaeological Investigations in Ancient America, edited by Wingard, John D. and Hayes, Sue E., pp. 85107. University Press of Colorado, Boulder.Google Scholar
Putnam, D. H., Oplinger, E. S., Doll, J. D., and Schulte, E. M. 1989 Amaranth. In Alternative Field Crops Manual, compiled by the University of Wisconsin Cooperative Extension Service, University of Minnesota Extension Service, and Center for Alternative Plant and Animal Products. Electronic document, https://hort.purdue.edu/newcrop/afcm/amaranth.html, accessed February 29, 2016.Google Scholar
Rand, Patricia J. 1965 Factors Related to the Distribution of Ponderosa and Pinyon Pines at Grand Canyon, Arizona. PhD dissertation, Department of Botany, Duke University, Durham, North Carolina.Google Scholar
Roberts, Neil 2011 Living with a Moving Target: Long-Term Climatic Variability and Environmental Risk in Dryland Regions. In Sustainable Lifeways: Cultural Persistence in an Ever-Changing Environment, edited by Miller, Naomi F., Moore, Katherine M., and Ryan, Kathleen, pp. 1338. University of Pennsylvania Museum of Archaeology and Anthropology, Philadelphia.Google Scholar
Rocek, Thomas R. 1995 Sedentarization and Agricultural Dependence: Perspectives from the Pithouse-to-Pueblo Transition in the American Southwest. American Antiquity 60:218239.Google Scholar
Roos, Christopher I. 2015 Western Apache Pyrogenic Placemaking in the Mountains of Eastern Arizona. In Engineering Mountain Landscapes: An Anthropology of Social Investment, edited by Scheiber, Laura L. and Zedeño, Maria N., pp. 116130. University of Utah Press, Salt Lake City.Google Scholar
Roos, Christopher I. 2017 Anthropogenic Landscapes. In Oxford Handbook of Southwest Archaeology, edited by Mills, Barbara and Fowles, Severin, pp. 683696. Oxford University Press, Oxford.Google Scholar
Roos, Christopher I., Scott, Andrew C., Belcher, Claire M., Chaloner, William G., Aylen, Jonathan, Bird, Rebecca B., Coughlan, Michael R., Johnson, Bart R., Johnston, Fay H., McMorrow, Julia, Steelman, Toddi, and the Fire and Mankind Discussion Group 2016 Living on a Flammable Planet: Interdisciplinary, Cross-Scalar, and Varied Cultural Lessons, Prospects, and Challenges. Philosophical Transactions of the Royal Society B 371: 20150469.Google Scholar
Roos, Christopher I., Sullivan, Alan P. III, and McNamee, Calla 2010 Paleoecological Evidence for Systematic Indigenous Burning in the Upland Southwest. In The Archaeology of Anthropogenic Environments, edited by Dean, Rebecca M., pp. 142171. Southern Illinois University Press, Carbondale.Google Scholar
Roos, Christopher I., and Swetnam, Thomas W. 2012 A 1416-Year Reconstruction of Annual, Multidecadal, and Centennial Variability in Area Burned for Ponderosa Pine Forests of the Southern Colorado Plateau Region, Southwest USA. Holocene 22:281290.Google Scholar
Ruddiman, William F. 2013 The Anthropocene. Annual Review of Earth and Planetary Science 41:4568.Google Scholar
Ruppé, Patricia A. 1985 A Long Pollen Record from Tsosie Shelter, Black Mesa, Arizona. In Excavations on Black Mesa, 1983: A Descriptive Report, edited by Christenson, Andrew L. and Perry, William J., pp. 513529. Center for Archaeological Investigations Research Paper No. 46. Southern Illinois University, Carbondale.Google Scholar
Salt Spring Seeds 2014 Growing Amaranth and Quinoa. Electronic document, https://www.saltspringseeds.com/pages/growing-amaranth-and-quinoa-dans-scoop, accessed June 18, 2018.Google Scholar
Scherjon, Fulco, Berkels, Corrie, MacDonald, Katharine, and Roebroeks, Wil 2015 Burning the Land: An Ethnographic Study of Off-Site Fire Use by Current and Historically Documented Foragers and Implications for the Interpretation of Past Fire Practices in the Landscape. Current Anthropology 56:299326.Google Scholar
Schwartz, Douglas W. 1990 On the Edge of Splendor: Exploring Grand Canyon's Human Past. School of American Research Press, Santa Fe, New Mexico.Google Scholar
Schwartz, Douglas W., Kepp, Jane, and Chapman, Richard C. 1981 Archaeology of the Grand Canyon: The Walhalla Plateau. School of American Research Press, Santa Fe, New Mexico.Google Scholar
Scott, Linda J. 1986 Analysis of Selected Macrofloral Samples from AZ I:1:17 (ASM). In Prehistory of the Upper Basin, Coconino County, Arizona, edited by Sullivan, Alan P. III, pp. 339346. Arizona State Museum Archaeological Series No. 167. University of Arizona, Tucson.Google Scholar
Smith, Bruce D. 2011 Shaping the Natural World: Patterns of Human Niche Construction by Small-Scale Societies in North America. In The Subsistence Economies of Indigenous North American Societies, edited by Smith, Bruce D., pp. 593609. Smithsonian Institution Scholarly Press, Washington, DC.Google Scholar
Smith, Bruce D. 2014 Documenting Human Niche Construction in the Archaeological Record. In Method and Theory in Paleoethnobotany, edited by Marston, John C., Guedes, Jade D'Alpoim, and Warinner, Christina, pp. 355370. University Press of Colorado, Boulder.Google Scholar
Smith, Bruce D., and Zeder, Melinda A. 2013 The Onset of the Anthropocene. Anthropocene 4:813.Google Scholar
Smith, Margaret 2017 Quinoa. Agricultural Marketing Resource Center, Iowa State University, Ames. Electronic document, https://www.agmrc.org/commodities-products/grains-oilseeds/quinoa/, accessed June 18, 2018.Google Scholar
Spielmann, Katherine A., Nelson, Margaret, Ingram, Scott, and Peeples, Matthew A. 2011 Sustainable Small-Scale Agriculture in Semi-Arid Environments. Ecology and Society 16. Electronic document, http://www.ecologyandsociety.org/vol16/iss1/art26, accessed April 20, 2014.Google Scholar
Springer, Judith D., Daniels, Mark L., and Nazaire, Mare 2009 Field Guide to Forest and Mountain Plants of Northern Arizona. Ecological Restoration Institute, Northern Arizona University, Flagstaff.Google Scholar
Stewart, Guy R., and Donnelly, Maurice 1943 Soil and Water Economy in the Pueblo Southwest. Scientific Monthly 56:134144.Google Scholar
Sullivan, Alan P. III 1996 Risk, Anthropogenic Environments, and Western Anasazi Subsistence. In Evolving Complexity and Environmental Risk in the Prehistoric Southwest, edited by Tainter, Joseph A. and Tainter, Bonnie B., pp. 145167. Santa Fe Institute Studies in the Sciences of Complexity Vol. XXIV. Addison-Wesley, Reading, Massachusetts.Google Scholar
Sullivan, Alan P. III 2000 Effects of Small-Scale Prehistoric Runoff Agriculture on Soil Fertility: The Developing Picture from Upland Terraces in the American Southwest. Geoarchaeology: An International Journal 15:291314.Google Scholar
Sullivan, Alan P. III 2008 Time Perspectivism and the Interpretive Potential of Palimpsests: Theoretical and Methodological Considerations of Assemblage Formation History and Contemporaneity. In Time in Archaeology: Time Perspectivism Revisited, edited by Holdaway, Simon and Wandsnider, LuAnn, pp. 3145. University of Utah Press, Salt Lake City.Google Scholar
Sullivan, Alan P. III 2015 The Archaeology of Ruderal Agriculture. In Traditional Arid Lands Agriculture: Understanding the Past for the Future, edited by Ingram, Scott E. and Hunt, Robert C., pp. 273305. University of Arizona Press, Tucson.Google Scholar
Sullivan, Alan P. III, Berkebile, Jean N., Forste, Kathleen M., and Washam, Ryan M. 2015 Disturbing Developments: An Archaeobotanical Perspective on Pinyon-Juniper Woodlands Fire Ecology, Economic Resource Production, and Ecosystem History. Journal of Ethnobiology 35:3759.Google Scholar
Sullivan, Alan P. III, and Forste, Kathleen M. 2014 Fire-Reliant Subsistence Economies and Anthropogenic Coniferous Ecosystems in the Pre-Columbian Northern American Southwest. Vegetation History and Archaeobotany 23:135151.Google Scholar
Sullivan, Alan P. III, and Ruter, Anthony H. 2006 The Effects of Environmental Fluctuations on Ancient Livelihoods: Implications of Paleoeconomic Evidence from the Upper Basin, Northern Arizona. In Environmental Change and Human Adaptation in the Ancient American Southwest, edited by Doyel, David E. and Dean, Jeffrey S., pp. 180203. University of Utah Press, Salt Lake City.Google Scholar
Swetnam, Thomas W., Allen, Craig D., and Betancourt, Julio L. 1999 Applied Historical Ecology: Using the Past to Manage the Future. Ecological Applications 9:11891206.Google Scholar
Thomas, David H. 1986 Refiguring Anthropology: First Principles of Probability and Statistics. Waveland Press, Prospect Heights, Illinois.Google Scholar
Uphus, Patrick M. 2003 The Influences of Ecosystem Variability on Prehistoric Settlement: Testing Terrain-Based Locational Models for the Upper Basin, Northern Arizona. Master's thesis, Department of Anthropology, University of Cincinnati, Cincinnati, Ohio.Google Scholar
Vankat, J. L. 2013 Vegetation Dynamics on the Mountains and Plateaus of the American Southwest. Springer-Verlag, Dordrecht, the Netherlands.Google Scholar
Van West, Carla R., and Lipe, William D. 1992 Modeling Prehistoric Climate and Agriculture in Southwestern Colorado. In The Sand Canyon Archaeological Project: A Progress Report, edited by Lipe, William D., pp. 105119. Occasional Paper No. 2. Crow Canyon Archaeological Center, Cortez, Colorado.Google Scholar
Wagner, Gail, Smart, Tristine, Ford, Richard I., and Trigg, Heather 1984 Ethnobotanical Recovery, 1982: Summary of Analysis and Frequency Tables. In Excavations on Black Mesa, 1982: A Descriptive Report, edited by Nichols, Deborah L. and Smiley, F. E., pp. 613632. Center for Archaeological Investigations Research Paper No. 39. Southern Illinois University, Carbondale.Google Scholar
West, Neil E. 1984 Successional Patterns and Productivity Potentials of Pinyon-Juniper Ecosystems. In Developing Strategies for Rangeland Management, pp. 13011332. National Research Council/National Academy of Sciences. Westview Press, Boulder, Colorado.Google Scholar
West, Neil E. 1999 Juniper-Pinyon Savanna and Woodlands of Western North America. In Savannas, Barrens, and Rock Outcrop Plant Communities of North America, edited by Anderson, Roger C., Fralish, James S., and Baskin, Jerry M., pp. 288308. Cambridge University Press, Cambridge.Google Scholar
Wilcox, David R. 1978 The Theoretical Significance of Fieldhouses. In Limited Activity and Occupation Sites, edited by Ward, Albert E., pp. 2532. Center for Anthropological Studies, Albuquerque, New Mexico.Google Scholar
Williams, Mark A., and Baker, William L. 2013 Variability of Historical Forest Structure and Fire across Ponderosa Pine Landscapes of the Coconino Plateau and South Rim of Grand Canyon National Park, Arizona, USA. Landscape Ecology 28:297310.Google Scholar
Wyckoff, Don G. 1977 Secondary Forest Succession Following Abandonment of Mesa Verde. Kiva 42:215231.Google Scholar
Yarnell, Richard A. 1965 Implications of Distinctive Flora on Pueblo Ruins. American Anthropologist 67:662674.Google Scholar
Figure 0

Table 1. Archaeological Studies that Mention Amaranth, Chenopodium, or “Cheno-ams” as Cultivated or Economically Significant Plants in the American Southwest.

Figure 1

Figure 1. Location of the Upper Basin, northern Arizona, showing excavated sites and area burned by the Scott Fire.

Figure 2

Figure 2. Agriculturally unproductive soils in the Upper Basin: (a) bedrock and very cobbly loam; (b) cobbly and very gravelly sandy loam; (c) very gravelly/sandy loam. (Color online)

Figure 3

Figure 3. Variation in the abundance and ubiquity of seeds and nuts (n = 3,485) identified in 110 samples recovered from features (postholes, thermal features, fire-cracked rock piles), artifacts (vessels, grinding stones), and occupation surfaces at 10 archaeological sites in the Upper Basin and Grand Canyon National Park (see Supplemental Table 1 for details). Each dot represents the frequency of nuts or seeds in a single sample, broken down by taxon (seeds classified by different archaeobotanists as cheno-am seeds, amaranth seeds, or chenopodium seeds are aggregated as “Cheno-am”). The data do not include counts of indirect indicators of plant use, such as cone scales, seed coats, nutshell, bark, needles, stems, leaves, wood, or cupules. This method was selected because it tightly constrains frequencies of edible plant parts—seeds or nuts—that in all likelihood were the objects of wild plant cultivation, wild plant gathering, or domesticated plant cultivation (Sullivan et al. 2015:44). Ubiquity values are given in parentheses.

Figure 4

Figure 4. Estimated time-corrected room counts, annual population, and hectares to be cultivated under different yields (kg/ha) of amaranth and chenopodium (based on data in Supplemental Table 2).

Figure 5

Table 2. Estimated Number of Fires Needed per Year Based on Different Productivity Estimates for Amaranth and Chenopodium.

Figure 6

Figure 5. Randomly placed, spatially scaled anthropogenic niches (1.1 km in diameter), with insert showing embedded burn plots (400 m in diameter) by time period in the Upper Basin.

Figure 7

Figure 6. Dynamics of anthropogenic niche and burn plot establishment and abandonment in the Upper Basin through time.

Figure 8

Figure 7. Upper Basin pinyon-juniper woodland (a) before (2008) and (b) after (2017) the lightning-caused Scott Fire (2016). (Color online)

Figure 9

Figure 8. Aftermath of the Scott Fire, which burned 1,076.5 ha in the Upper Basin between June 28 and July 18, 2016, showing dead trees and prehistoric masonry structure surrounded by fetid goosefoot in April 2017 (image used with permission of Neil Weintraub, Kaibab National Forest, US Department of Agriculture Forest Service).

Figure 10

Table 3. Forest Fires in the Upland American Southwest that Produced Amaranth or Chenopodium.

Supplementary material: File

Sullivan and Mink supplementary material

Sullivan and Mink supplementary material 1

Download Sullivan and Mink supplementary material(File)
File 86.2 KB