8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of...

30
142 8. Paleoecological Evidence for Systematic Indigenous Burning in the Upland Southwest Christopher I. Roos, Alan P. Sullivan III, and Calla McNamee Abstract: In the fire-prone coniferous ecosystems of the American South- west where archaeological sites are abundant and the ecological legacies of ancient societies are incompletely understood, the “anthropogenic fire” hypothesis has been underappreciated. We use a multiproxy, stratigraph- ic approach to reconstruct variation in fire regimes over the periods of peak prehistoric population in two areas of the upland Southwest. Com- parisons of these stratigraphic records to a tree-ring-based climate recon- struction indicate that the prehistoric residents of the Grand Canyon ( a.d. 850–1150/1200) and late prehistoric ( a.d. 1200–1400) and protohistoric ( a.d. 1600–1870) residents of the Mogollon Rim region used fire in different ways as part of food production strategies. These analyses indicate that aboriginal landscape management practices, which included systematic low-intensity burning, generated upland ecosystems markedly different from those en- countered today. As the contributions to this volume illustrate, the investigation of human impacts on past environments has gathered momentum within archae- ology. Increasing attention to human-modified environments has led some in- vestigators to conclude that anthropogenic environments are ubiquitous, even in the precontact Americas (e.g., Denevan 1992; Mann 2005). However, the ubiq- uity of culturally modified ecosystems is not universally accepted. Particularly with respect to North America, scientists outside of anthropology and human The Archaeology of Anthropogenic Environments, edited by Rebecca M. Dean. Center for Archaeo- logical Investigations, Occasional Paper No. 37. © 2009 by the Board of Trustees, Southern Il- linois University. All rights reserved. ISBN 978-0-88104-094-4.

Transcript of 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of...

Page 1: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

142

8. Paleoecological Evidence for Systematic Indigenous Burning in the Upland Southwest

Christopher I. Roos, Alan P. Sullivan III, and Calla McNamee

Abstract: In the fire-prone coniferous ecosystems of the American South-west where archaeological sites are abundant and the ecological legacies of ancient societies are incompletely understood, the “anthropogenic fire” hypothesis has been underappreciated. We use a multiproxy, stratigraph-ic approach to reconstruct variation in fire regimes over the periods of peak prehistoric population in two areas of the upland Southwest. Com-parisons of these stratigraphic records to a tree-ring-based climate recon-struction indicate that the prehistoric residents of the Grand Canyon (a.d. 850–1150/1200) and late prehistoric (a.d. 1200–1400) and protohistoric (a.d. 1600–1870) residents of the Mogollon Rim region used fire in different ways as part of food production strategies. These analyses indicate that aboriginal landscape management practices, which included systematic low-intensity burning, generated upland ecosystems markedly different from those en-countered today.

As the contributions to this volume illustrate, the investigation of human impacts on past environments has gathered momentum within archae-ology. Increasing attention to human-modified environments has led some in-vestigators to conclude that anthropogenic environments are ubiquitous, even in the precontact Americas (e.g., Denevan 1992; Mann 2005). However, the ubiq-uity of culturally modified ecosystems is not universally accepted. Particularly with respect to North America, scientists outside of anthropology and human The Archaeology of Anthropogenic Environments, edited by Rebecca M. Dean. Center for Archaeo-logical Investigations, Occasional Paper No. 37. © 2009 by the Board of Trustees, Southern Il-linois University. All rights reserved. ISBN 978-0-88104-094-4.

Page 2: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

Systematic Indigenous Burning in the Upland Southwest 143

geography are less willing to accept arguments that precontact environments were affected by human land use in any significant way (e.g., Vale 2002). These scientists emphasize empirical studies of past environments often with the ex-plicit purpose of using paleoenvironmental data to identify the local conse-quences of climate changes. Human-induced environmental changes, if pres-ent, are often considered by them to be “noise” because climate is assumed to be the primary driver of “natural” environmental change over the long term (cf. Burroughs 2005). If archaeologists are genuinely interested in contributing to the dialog on human- and climate-related environmental changes (e.g., Redman et al. 2004), then we must engage “natural” scientists on their terms. This strategy means that the ubiquity of anthropogenic environments must not be dogmatically as-sumed. To persuade natural scientists that anthropogenic environments are a significant part of North American ecological histories, therefore, we must as-semble relevant evidence that can be used to evaluate the null hypothesis that climate was the sole driver of environmental change. This challenge is daunting because identifying anthropogenic environmental signals may not be possible by employing traditional archaeological or paleoecological approaches (Dean 2004). The foregoing argument is particularly salient for anthropologists interested in the cultural manipulation of environments through the use of fire (Stewart 2002). Ethnographers and ethnohistorians have documented nearly ubiquitous landscape burning by American Indians (e.g., Pyne 2000). Aboriginal uses of fire in ecosystem management are plentiful and include descriptions of its role in hunting, agriculture, wild-plant productivity, pest control, and warfare (Williams 2002). Many fire historians, however, particularly those who are familiar with fire-prone areas in the American Southwest (Allen 2002), do not readily accept the suggestion that American landscapes were widely shaped by Indian burning (contributions in Vale, ed. 2002). Yet, from an anthropological perspective, hypotheses concerning the use of landscape burning to manage wild-plant predictability and productivity (Sul-livan 1996) have implications for understanding variation in prehistoric liveli-hoods that became widespread after the introduction of maize approximately 3,500 years ago (Bonnicksen 2000:113–119). From a stewardship perspective, understanding the presence or absence of systematic indigenous burning, its ecological consequences, and resilience to climatic or land use changes have implications for deciding how ecosystem management and restoration should be structured today and in the future (Bonnicksen et al. 1999; Morgan et al. 2003). Nonetheless, the identification of human-influenced fire regimes will not be convincing unless other causal factors that affect variability in fire histo-ries and their ecological consequences are considered (Periman et al. 1999). For these reasons, we investigate the relationships among climate, fire, and land use histories for two study areas in the upland Southwest (Figure 8-1) within a null model framework (Connor and Simberloff 1986).This approach allows us to evaluate the hypothesis that fire-regime variation was attributable exclusively to climatic change.

Page 3: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

144 C. I. Roos, A. P. Sullivan III, and C. McNamee

Presettlement Fire Regimes of the Upland Southwest

According to most scholars, prior to extensive grazing and active fire suppression, ponderosa pine (Pinus ponderosa) forests experienced low-intensity fires every 3–10 years that maintained an open-canopied, parklike forest (e.g., Fulé et al. 1997). The occurrence and extent of fires in a given year was depen-dent upon the properties and distribution of fuels. Although ponderosa pine forests “manage” their own fuel bed by rapidly dehiscing needles (DeBano et al. 1998:2–3), fires during the historic period (a.d. 1700–1900) were facilitated by the production of fine, herbaceous understory fuels during multiple wet years that were cured during subsequent dry years (Crimmins and Comrie 2004; Swet-nam and Baisan 1996, 2003). Ponderosa pine trees have a number of fire-adapted traits, including thick bark, self-pruning branches, and high crown-scorch toler-ance, that imply ponderosa pine forests evolved in the context of high-frequency surface fires (Covington 2003). The historical role of fire in the development of woodlands dominated by pin-yon (Pinus edulis or Pinus cembroides) and juniper (Juniperus utahensis or Juniperus monosperma) is a controversial topic (Baker and Shinneman 2004). It has been al-ternately suggested that pinyon pine, in particular, is very sensitive to fire or very

Figure 8-1. Location of the Upper Basin and Forestdale Valley study areas in relation to circum–Colorado Plateau coniferous environments. Dark gray ar-eas are ponderosa pine forests. Light gray areas are juniper or pinyon-juniper woodlands.

FPO

Page 4: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

Systematic Indigenous Burning in the Upland Southwest 145

difficult to kill with fire (cf. Floyd et al. 2003:264; Pieper and Wittie 1990; Wittie and McDaniel 1990:176). Reconstructions of historic-period fire regimes in pinyon-juniper forests have been difficult because the major species rarely record fire scars. Stand-age studies in the northern Southwest have suggested that fires in pinyon-juniper were probably high-severity crown fires but were exceedingly rare over the past few hundred years (Floyd et al. 2004; Floyd et al. 2003; Floyd et al. 2008). In the few studies that have reported fire-scar records for pinyon, surface fire return intervals have ranged from 5 to 74 years in the Texas borderlands (Poulos 2007:107) and 10 to 49 years in southern New Mexico (Brown et al. 2001:121). The emerging picture of pinyon-juniper fire ecology is one of variability, in part dependent upon the density, composition, and structure of understory and canopy species (Romme et al. 2008). Beyond the past 500 years, however, the fire ecology and fire history of pinyon-juniper woodlands, shrublands, and savannahs are not well known. In the Madrean province of southern Arizona, southern New Mexico, and northern Mexico, pinyon-juniper woodland on the United States side of the inter-national border is more dense with a discontinuous, shrubby understory in com-parison to pinyon-juniper woodland on the Mexican side of the border, which has been subject to less intensive fire suppression (Kruse et al. 1996:102). This pat-tern suggests that prior to extensive grazing and fire suppression, some pinyon-juniper woodlands may have been more open with a diverse, herbaceous under-story (Barney and Frischknecht 1974:91). From a fire-regime perspective, such an understory composition would have been significant because fires in grassy fuels are often faster moving and lower in intensity than shrub fires (DeBano et al. 1998:59–61). Consequently, fires in an open canopy and grassy understory may be less likely to induce high mortality rates in mature pinyon or juniper trees (Romme et al. 2008:7).

Study Areas

To evaluate the consequences of anthropogenic fire in Southwestern coniferous uplands, we chose two study locales whose rich archaeological heri-tages have been the subject of hypotheses concerning prehistoric burning—the Upper Basin of northern Arizona and the Forestdale Valley of east-central Arizo-na (Figure 8-1). These two study areas allow us to contrast fire regimes between pinyon-juniper woodlands (Upper Basin) and ponderosa pine forests (Forestdale Valley), as well as to evaluate the role of burning for wild-plant management (Upper Basin) and agricultural production (Forestdale Valley) during Ancestral Pueblo occupations.

The Upper Basin The Upper Basin is a graben of the northeastern Coconino Plateau. El-evations in the Upper Basin range from 2,195 m at Desert View on the South Rim of the Grand Canyon to 1,829 m at the bottom of Lee Canyon. Surficial geology is almost exclusively Kaibab Formation limestone and sandstone (Hopkins and

Page 5: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

146 C. I. Roos, A. P. Sullivan III, and C. McNamee

Thompson 2003). Vegetation of the Upper Basin consists primarily of pinyon-juniper woodlands interspersed with pockets of sagebrush-grasslands (Brewer et al. 1991). Although the Upper Basin experiences at least 130 consecutive frost-free days and receives an average of 36–42 cm of precipitation annually (Hendricks 1985:127; Merkle 1952:376; Sullivan and Ruter 2006:185–187), a pronounced sea-sonal drought characterizes the late spring and early summer, as well as the early fall (Sullivan et al. 2002:53). More than 20 km2 of the Upper Basin, in the Kaibab National Forest, have been intensively surveyed (inter-surveyor distance = 10 m) by the Upper Basin Archaeological Research Project (UBARP). This survey has recorded the location and surface expression of 250 masonry ruins, 148 fire-cracked rock piles, and 790 artifact scatters, as well as other features (Sullivan et al. 2007). The vast majority of these remains date to the primary occupation of the Grand Canyon region, a.d. 850–1150/1200. On the basis of well-preserved paleobotanical assemblages recovered from catastrophically abandoned settlements (Sullivan 1987) and extra-mural plant-processing locations (Sullivan 1992), the prehistoric residents of the Upper Basin practiced a mixed gathering-gardening subsistence strategy that em-phasized wild-plant resources (Sullivan 1986), which may have been propagated by controlled burning of understory vegetation (Sullivan 1996). By the early thir-teenth century, Ancestral Pueblo people had ceased to occupy the Upper Basin on a perennial basis. Episodic use of the area thereafter by Hopi, Pai, and ultimately Navajo groups occurred sporadically during the protohistoric and historic peri-ods (Cooper 1960:138; Morehead 1996; Schwartz 1990; Sullivan et al. 2001). In 2000, UBARP placed backhoe trenches across Simkins Flat, which is an alluvial channel fan in Red Hawk Wash. Exposed sediments in the trenches were sampled for geoarchaeological and paleoecological analyses in 2000, 2001, and 2002 (McNamee 2003; Sullivan and Ruter 2006). Data from these original inves-tigations, as well as the results of new geochemical analyses, are used in this study.

The Forestdale Valley The Forestdale Valley is an upland alluvial valley excavated primar-ily into Cenozoic- and Permian-aged sedimentary rocks (Haury 1985 [1940]:142). Forestdale Creek drains south–southwest from headwaters along the Mogollon Rim, a fault scarp that defines the southern edge of the Colorado Plateau, to a Tertiary basalt flow that affects local base level and is inferred to be responsible for the unusually thick alluvial deposits in the valley (Haury 1985:13). Located just south of the modern town of Show Low, the Forestdale Valley is situated on the Fort Apache Indian Reservation near the center of the largest stand of ponde-rosa pine forest in the Southwest (Covington 2003). Elevations in the valley range from 1,791 m at the confluence with Corduroy Creek to 2,105 m along the Mog-ollon Rim. The valley experiences an average of 49 cm of annual precipitation split bimodally between summer rainfall and winter rain and snowfall. Because of cold air drainage, the valley experiences an average of 108 frost-free days (Ka-ldahl and Dean 1999:13).

Page 6: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

Systematic Indigenous Burning in the Upland Southwest 147

The eastern Mogollon Rim region (defined here as the coniferous belt from Chevelon Creek on the west to the White Mountains on the east), which includes the Forestdale Valley, has been the subject of archaeological research for more than a century (Haury 1985; Hough 1903; Mills et al. 1999; Reid and Whittlesey 1999). Although humans may have occupied the region since the Terminal Pleisto-cene (ca. 13,000 cal b.p.), there is very little archaeological evidence for human use of the Forestdale area until the Pithouse Period (after a.d. 200). Pithouse Period archaeology (a.d. 200–1000) is characterized by dispersed, seasonally occupied small settlements (Gilman 1997; Haury 1985 [1940]; Haury and Sayles 1985 [1947]). Pithouse occupants practiced limited horticulture (Geib and Huckell 1994) but relied heavily on wild plant and animal resources, which probably resulted in an extensive rather than intensive environmental influence. Although there was a major population influx shortly after a.d. 1000 (Herr 2001; Newcomb 1999), oc-cupations remained short-term (subdecadal to decadal) and land use continued to involve a mixed exploitation of wild and cultivated resources (Huckell 1999; Welch 1996). By a.d. 1300, most of the area’s residents aggregated into a handful of large, perennially occupied villages (Kaldahl et al. 2004), resulting in a more intensive land use strategy that was heavily reliant upon agriculture (Ezzo 1992; Welch 1996). By the end of the fifteenth century, the Western Pueblo people ceased to occupy the area on a perennial basis and moved their permanent residences else-where (Mills 1998). The use of fire may have been important in Pueblo agricul-tural strategies, such as in “burn-plot” style agriculture (Sullivan 1982) or, less likely, in swidden-style field clearance (Kohler and Mathews 1988). Early Apache occupation of the area is poorly understood, but Spanish doc-uments indicate the presence of horticultural Apaches in the Upper Gila area by a.d. 1626 (Buskirk 1986:109). This report may be supported archaeologically by a terminus ante quem date of a.d. 1661 from Wickiup 2 at the Grasshopper Spring site (Whittlesey et al. 1997:198). According to oral tradition, Western Apaches used fire as an important part of hunting (Buskirk 1986:127, 131, 135–136), gath-ering (Buskirk 1986:165–166; Gifford 1940), and horticultural subsistence prac-tices (Buskirk 1986:43, 61, 77). In 2005, the Mogollon Rim Historical Ecology Project (MRHEP) initiated geoarchaeological and paleoecological fieldwork along the eastern Mogollon Rim. In collaboration with the White Mountain Apache Tribal Heritage Program, three Terrace 3 aged sedimentary deposits (ca. a.d. 1150–1910 according to An-tevs in Haury 1985 [1940]:143–144) from the Forestdale Valley, upstream from most archaeological sites, were sampled by MRHEP (Roos 2008). Data from two localities, Forestdale Valley 6 and 10, are used here.

Discussion For both the Upper Basin and the Forestdale Valley study areas, chron-ological control is provided by radiocarbon dates on detrital charcoal extracted from alluvial sediments and soils (Table 8-1) derived from Bayesian calibration (Buck et al. 1991) with the Intcal04 data set (Reimer et al. 2004) in the BCal online

Page 7: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

148 C. I. Roos, A. P. Sullivan III, and C. McNamee

calibration software (Buck et al. 1999; http://bcal.sheffield.ac.uk). The calibrated age ranges are reported in calendar years a.d. at the 95-percent confidence in-terval. For reference, the traditional 2σ calibrated ages (using CALIB 5.0 and the Intcal04 data set) are reported in Table 8-1 as well. Trench 1 on Simkins Flat disclosed evidence of sediment accumulation be-tween cal a.d. 779–973 and the present, punctuated by episodes of erosion and braided channel formation between radiocarbon ages of cal a.d. 1045–1215 and cal a.d. 1416–1503. Archaeological and radiocarbon data indicate that sediments accumulated at Forestdale 6 and 10 between a.d. 1100 and a.d. 1910 (Haury 1985:145; Roos 2008:176–178). Although there is no evidence for erosion in either of these sequences, a brief episode of soil stability and channel entrenchment may have occurred in upper Forestdale Creek between a.d. 1400 and a.d. 1500 prior to resumed aggradation (Roos 2008:181).

Long-Term Variation in Fire-Climate Conditions

Many fire-history studies in the U.S. Southwest have evaluated fire-climate relations (e.g., Grissino-Mayer and Swetnam 2000; Kitzberger et al. 2001; Swetnam and Baisan 1996; Swetnam and Betancourt 1990, 1998). Most of these studies compare annually dated fire scars (Arno and Sneck 1977; Dieterich and Swetnam 1984) on living trees, snags, and stumps to interannual (Swetnam and Baisan 1996, 2003), decadal (Swetnam and Betancourt 1998), and centennial-scale (Grissino-Mayer and Swetnam 2000) climate-change data that are reconstructed from independent dendroclimatic models. Fire-scar–based fire histories in the U.S. Southwest are limited by an age-attenuation effect (Clark 1988:81; Whitlock and Bartlein 2004:479) to the past 300–400 years. Importantly, because fire-scar records are too short to rule out legacies of prehistoric human impacts to historic fire regimes, when these records have been used to evaluate human influences on fires and their frequencies, results often have been equivocal (e.g., Allen 2004:51–54; Grissino-Mayer et al. 2004). The availability of annually resolved fire scars, the abundance of annually resolved local dendroclimate records from the Southwest Paleoclimate Project (Dean and Robinson 1977), and the statistical association of interannual moisture and fires (Crimmins and Comrie 2004; Swetnam and Baisan 1996, 2003), how-ever, create an ideal situation for generating long-term reconstructions of fire-climate conditions that are independent of long-term fire-history reconstructions. For instance, Roos and Swetnam (2009) developed a multiple regression model of antecedent moisture conditions from dendroclimatic reconstructions based on 46 ponderosa pine fire-scar chronologies and two regional precipitation reconstruc-tions (Grissino-Mayer 1996; Salzer and Kipfmueller 2005) across the southern margins of the Colorado Plateau. This model explains approximately 38 percent of the variation in landscape fire size (r2 = .381, p < .0001) during the historic period (a.d. 1700–1899). In addition, it characterizes long-term variability in the frequency of regional fire-climate years by generating a 100-year moving average of a binary variable created from the presence or absence of peak fire-climate con-

Page 8: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

Systematic Indigenous Burning in the Upland Southwest 149

Tabl

e 8-

1. S

imki

ns F

lat a

nd F

ores

tdal

e Val

ley 6

and

10 R

adio

carb

on A

ges

Sam

ple #

Dep

thM

ater

ial

14C

2σ C

ala

95%

CI C

alb

Sim

kins

Fla

t

RC-5

53U

ID c

harc

oal

610

± 50

1286

–141

3 (1

.00)

1465

–165

1RC

-183

Out

er ri

ng fr

om b

urne

d Pi

nus e

dulis

log

425

± 40

1574

–162

6(.1

5)14

17–1

522(

.85)

1416

–150

3

RC-4

100

UID

cha

rcoa

l90

0 ±

3810

36–1

213

(1.0

0)10

45–1

215

RC-2

142

UID

cha

rcoa

l11

41 ±

36

804

–984

(.96

) 7

80–7

92(.0

4)80

1–97

377

9–79

4Fo

rest

dale

6

6.2.

2447

UN

WC

212

± 54

1909

–195

3 (.1

5)17

16–1

891

(.51)

1629

–171

1 (.2

8)15

22–1

573

(.06)

1632

–181

6

6.5

75Tw

ig39

0 ±

3315

58–1

631

(.31)

1440

–152

4 (.6

9)15

55–1

640

6.11

143

Con

ifer x

ylem

355

± 38

1536

–163

5 (.5

4)14

53–1

533

(.46)

1453

–158

2

6.5.

8817

5Pi

nus n

eedl

e62

9 ±

9512

18–1

447

(1.0

0)13

91–1

504

6.6.

119

237

Pinu

s mer

iste

m47

7 ±

6915

58–1

631

(.12)

1384

–152

4 (.7

4)13

04–1

365

(.14)

1378

–145

813

32–1

356

6.8.

150

299

Pinu

s mer

iste

m77

7 ±

6513

61–1

386

(.03)

1150

–130

9 (.9

1)11

23–1

138

(.01)

1048

–108

6 (.0

4)

1339

–139

513

03–1

324

Page 9: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

150 C. I. Roos, A. P. Sullivan III, and C. McNameeTa

ble

8-1.

—Co

ntin

ued

Sam

ple #

Dep

thM

ater

ial

14C

2σ C

ala

95%

CI C

alb

6.26

341

Con

ifer x

ylem

535

± 57

1377

–144

9 (.5

7)12

98–1

372

(.43)

1286

–135

6

6.29

369

Con

ifer x

ylem

973

± 59

972

–121

1 (1

.00)

966–

1176

Fore

stda

le 1

0

10.1

.17

33St

em37

1 ±

3415

44–1

634

(.44)

1446

–152

9 (.5

6)15

51–1

641

1488

–152

510

.711

8Pi

nus n

eedl

e61

5 ±

3412

92–1

403

(1.0

0)13

61–1

423

10.4

.63

125

UN

WC

518

± 64

1292

–148

3 (1

.00)

1326

–144

110

.10

159

Twig

439

± 34

1601

–161

5 (.0

3)14

14–1

500

(.96)

1418

–148

6

10.5

.92

183

UN

WC

619

± 38

1289

–140

4 (1

.00)

1292

–136

610

.16

247

UN

WC

672

± 35

1347

–139

2 (.4

4)12

71–1

323

(.56)

1266

–131

1

10.7

.138

275

Pinu

s nee

dle

1120

± 1

2010

64–1

155

(.09)

668

–105

9 (.9

1)10

56–1

257

10.8

.152

303

Pinu

s nee

dle

800

± 97

1024

–131

6 (.9

5)96

5–11

44

Not

e: A

ll da

tes

wer

e m

easu

red

usin

g A

MS

with

the

exce

ptio

n of

RC

-1 a

nd R

C-5

, whi

ch w

ere

mea

sure

d ra

diom

etric

ally

. All

calib

rate

d ag

es a

re a

.d. u

nles

s ot

herw

ise

note

d. U

ID c

harc

oal =

uni

dent

ified

cha

rcoa

l; U

NW

C =

uni

dent

ified

non

-woo

d ch

arco

al.

Sour

ce: S

imki

ns F

lat d

ata

adap

ted

from

McN

amee

200

3; F

ores

tdal

e Va

lley

6 an

d 10

dat

a ad

apte

d fr

om R

oos 2

008.

a Cal

ibra

ted

in C

ALI

B 5.

01 u

sing

Intc

al 2

004.

Num

bers

in p

aren

thes

es in

dica

te th

e re

lativ

e ar

ea u

nder

the

curv

e (in

%) f

or th

e re

spec

tive

calib

rate

d ag

e ra

nge.

b Baye

sian

cal

ibra

tion

func

tion:

Si

mki

ns F

lat:

RC–5

> R

C–1

> R

C–4

> R

C–2

.

Fore

stda

le 6

: a.d

. 191

0 >

6.2.

24 ≥

6.5

> 6

.11

> 6.

5.88

> 6

.6.1

19 >

6.8

.150

> 6

.26

> 6.

29.

Fo

rest

dale

10:

a.d

. 191

0 >

10.1

.17

> 10

.7 >

10.

4.63

> 1

0.10

> 1

0.5.

92 >

10.

16 >

10.

7.13

8 >

10.8

.152

.

Page 10: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

Systematic Indigenous Burning in the Upland Southwest 151

ditions (≥.5 standard deviations above the long-term mean). Figure 8-2 plots both annual model values and long-term frequencies of fire-conducive climate condi-tions in Z-scores (standard deviation units about the long-term mean).Predicted frequencies of fire-conducive moisture patterns were below average (less than 31 events per century) in the two periods a.d. 650–750 and a.d. 1300–1650, which, under exclusively “natural” fire regimes, would have resulted in less frequent fires, the accumulation of fuels, and the succession of understory plant commu-nities to more shrub-dominated arrangements. In contrast, fire-climate frequen-cies were above average (more than 31 events per century) in the periods a.d. 750–850, a.d. 1100–1300, and after a.d. 1650. Although this model was generated from ponderosa pine fire-scar chronologies, the patterns of interannual moisture are probably relevant for surface fuels in all woodland environments, including pinyon-juniper (Crimmins and Comrie 2004).

Reconstructed Fire Regimes

As noted above, some coniferous environments in the uplands of the U.S. Southwest were characterized by high fire frequencies prior to modern fire suppression and grazing. Tracking changes in fire frequency alone, therefore, may not be a very useful measure of the extent of anthropogenic burning. The concept of the fire regime, which includes consideration of the frequency, inten-sity, seasonality, average size, type of fuels, and overall ecological impact of land-scape burning, provides a better framework for investigating anthropogenic fire history. Therefore, we use multiple lines of evidence to reconstruct fire regimes, including sedimentary charcoal, grain size, soil phosphorus, stable-carbon iso-tope, and stratigraphic palynological data from alluvial channel fan sequences within the two study areas (cf. Adams 2004). These data will allow us to infer fire size and frequency (charcoal and soil phosphorus [P]), seasonality (grain size and carbon isotopes), fuel type (charcoal), and ecological impact (pollen and carbon isotopes) of past fires (Table 8-2). Although sedimentary charcoal analysis is most often conducted on samples from more or less constantly aggrading and uneroded sedimentary sequences, such as lakes and wetlands (Whitlock and Anderson 2003), these types of depo-sitional contexts are rare in the U.S. Southwest. In contrast, alluvial sedimentary contexts are abundant and often of direct relevance to locations of prehistoric and protohistoric occupation (Dean et al. 1985; Karlstrom 1988). Two properties of alluvial channel fans from small discontinuous ephemeral streams in small watersheds (<40 km2) make them well suited for investigating anthropogenic fire: (1) relatively rapid and steady aggradation punctuated by brief episodes of entrenchment (Bull 1997; Huckleberry and Billman 1998) and (2) local, slope-derived sources of sediment (Balling and Wells 1990; Hereford 2002). Moreover, experimental studies have documented the reliability of macroscopic sedimenta-ry charcoal (>125 μm) for identifying local, watershed-scale fires (Blackford 2000; Clark et al. 1998; Clark and Royall 1995; Lynch et al. 2004; Whitlock and Mills-paugh 1996; Whitlock et al. 2004).1 Although variation in “background” charcoal

Page 11: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

152 C. I. Roos, A. P. Sullivan III, and C. McNamee

Figure 8-2. Fire-conducive antecedent moisture patterns for the southern Colorado Plateau between a.d. 700 and a.d. 1900 (adapted from Roos and Swetnam 2009). Vertical bars indicate annual reconstruction values. The dark line indicates long-term frequencies of peak (≥.5 Z units) fire-climate conditions as 100-year moving averages of “peak” occurrence. Both variables are plotted in Z-scores about the long-term mean for each record.

(unrelated to fire frequencies) has been interpreted in terms of changes in the type of biomass burned (Marlon et al. 2006), most studies interpret variation in charcoal in terms of the amount of biomass consumed (fire size). This may be an appropriate protocol for fire regimes characterized by woody understory vegeta-tion or closed canopies and fire-return intervals measured in decades or centuries. However, in high-frequency surface fire regimes, such as those that characterize ponderosa pine forests and have been hypothesized for some pinyon-juniper woodlands (Romme et al. 2008; Sullivan 1996), changes in the type of biomass combusted may have an equally significant effect on sedimentary charcoal abun-dance (e.g., Marlon et al. 2006). Herbaceous plants produce fewer particulate by-products than woody plants, in part due to increased efficiency of combustion (DeBano et al. 1998:27). Consequently, a fire regime in which surface fires occur frequently enough to maintain herbaceous dominated understory plants (<5–10 years per plot) and burn predominantly within this vegetation would produce far less charcoal than a fire regime burning over similar-sized areas but within shrubby vegetation. Wildland fire accelerates the mineralization of organically bound phosphorus (P) and often results in net increases in soil phosphorus after low- or moderate-

FPO

Page 12: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

Systematic Indigenous Burning in the Upland Southwest 153

Table 8-2. Geoarchaeological Proxies and Number of Samples Used to Reconstruct Fire Regimes

Proxy

Sample N (Simkins Flat/

FDV 6/FDV 10) Expectations

Sedimentary charcoal

66/179/161 Increased abundance =

Decreased abundance =

More frequent burning of woody fuels OR larger fire sizes in woody fuelsLess frequent burning of woody fuels OR increase in patch-specific fire-return intervals (grassy fuels)

Soil P 8/25/15 Increasing concentrations =Decreasing concentrations =

Increasing fire frequencyDecreasing fire frequency

δ13C 8/25/15 Less negative values =

More negative values =

Increasing abundance of C4 plants (frequent patch-specific burning)Decreasing abundance of C4 plants (less-frequent patch-specific burning)

Palynology 11/15/10 Higher representation of herbaceous taxa =

Higher representation of woody taxa =

More frequent patch-specific burning

Less frequent patch-specific burning

Note: FDV = Forestdale Valley.

intensity fires (Covington and DeBano 1990; Covington and Sackett 1990; DeBano et al. 1998:120). Phosphorus that is mineralized from organic forms during com-bustion simultaneously becomes available to plants as well as erosion. If a portion of these phosphate minerals is deposited in alluvial contexts or if fuels on alluvial channel fans burn at the same time as surrounding hillslopes, stratigraphic soil phosphorus from alluvial deposits should provide an independent measure of bio-mass burning (and organic phosphorus conversions to mineral phosphorus). Stable carbon isotope ratios have been used to track changes in the contribu-tions of C3 and C4 plants to soil organic carbon pools (Biedenbender et al. 2004; Nordt 2001) and, therefore, can be used to assess the role of anthropogenic fire in maintaining disturbance taxa (e.g., Bowman et al. 2004). C4 plants, with few excep-tions, are “warm-season” herbaceous plants (grasses, weeds, forbs) that fractionate less against the 13C isotope and produce stable carbon isotope ratios (δ13C) in the range of about –12‰. Many important economic plants to ancient Southwestern-

Page 13: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

154 C. I. Roos, A. P. Sullivan III, and C. McNamee

ers, such as chenopodium, amaranth, and maize, are C4 plants (Wills 1995:229). C3 plants, which include some herbaceous plants but also virtually all woody shrubs and trees, fractionate more against 13C and produce δ13C values around –25‰. Stable carbon isotope ratios of soil organic matter are representative of the plants that have grown and decayed there and, as a result, produce a mixed value that depends upon the proportion of C3 and C4 plants contributing to decayed organic matter. With high-frequency burning, which promotes the growth of herbaceous vegetation, δ13C values of upland soils should be expected to increase (become less negative). Less-negative δ13C values are indicative of increasing proportions of C4 plants on the landscape (either on the fan, surrounding slopes, or both). Seasonality of burning may complicate the relationship between fire frequencies and carbon isotope ratios. For example, autumn burning promotes the growth of perennial cool-season grasses, which are C3 plants. In contrast to carbon isotopes of soil organic carbon, which can lag veg-etation changes, pollen assemblages respond immediately to changes in plant composition and are sensitive to more than two taxa (e.g., C3 vs. C4). Variation in sequences of pollen taxa assemblages provides a more direct measurement of vegetation change related to climate, disturbance, or land use (e.g., Wyckoff 1977). Pollen taxa were aggregated into three categories for discussion here: trees, shrubs, and herbs.2 Although varying proportions of these taxa do not directly re-late to their abundance on the surrounding landscape, the relative abundance of these three pollen taxa is relevant to the investigation of fire history. Herbaceous plants, including grasses, weeds, and forbs, dominate the understory of forests and woodlands that burn with a high frequency. Additionally, these plants in-clude many wild plants of economic importance to prehistoric and protohistoric residents of the Mogollon Rim and Grand Canyon areas (Buskirk 1986; Sullivan 1987, 1992; Sullivan and Ruter 2006).

Upland Fire-Regime Variability

The data from Simkins Flat (Figure 8-3) produce an integrated pic-ture of local fire-regime variation for the Red Hawk Wash watershed. Prior to a.d. 1100, high soil phosphorus concentrations, low charcoal concentrations, and elevated representation of herbaceous pollen taxa indicate very high-frequency fires that burned almost exclusively within herbaceous vegetation (low char-coal concentrations), and which may have actively promoted herbaceous plant production on the channel fan itself. Despite the low representation of arboreal taxa in pollen assemblages, pinyon and juniper trees must have been relative-ly abundant away from the fan. In fact, contemporaneous archaeological sites above the fan yield abundant arboreal pollen (see Table 8-3), which suggests that the reduced arboreal pollen representation in the Simkins Flat sequence is due to localized abundance of herbaceous plants.The large charcoal peak (burning in woody fuels) that immediately precedes an erosion event may coincide with archaeological evidence for catastrophic burning of at least one settlement after a.d. 1064 (Sullivan 1987) and exceptional fire-climate conditions in a.d. 1067 (see

Page 14: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

Systematic Indigenous Burning in the Upland Southwest 155

Figure 8-3. Stratigraphic data for Trench 1 on Simkins Flat in the Upper Basin. Approximate occupation periods are marked on the right (see also Mc-Namee 2003; Sullivan and Ruter 2006:200). Vertical dashed lines indicate overall mean values for grain size, charcoal, soil P, and δ13C.

Table 8-3. Arboreal Pollen Percentages for Archaeological Sites and Alluvial Sediments from the Upper Basin

Site Type of Features Dates a.d.

Range of Arboreal Pollen Relative

Abundances (%)

MU 125 Multiroom masonry structure 1070–1080 30–75

MU 235 Fire-cracked-rock pile417–559687–963 55–80

MU 125A Stone alignments and terraces ca. 1050–1150 25–80

ASM 17 Multiroom masonry and wood structures 1049/1054–1064 46–75

Simkins Flat Alluvial sedimentsca. 800–1200ca. 1200–1900

15–3650–75

Note: Arboreal pollen abundances between 50 and 80 percent are common for “natural” conifer wood-land settings whereas abundances below 50 percent occur in the context of local herbaceous plant production but do not imply canopy reduction. Contemporaneous samples from less than 2 km away from Simkins Flat indicate “natural” woodland conditions, even in proximity to archaeological sites.

FPO

Page 15: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

156 C. I. Roos, A. P. Sullivan III, and C. McNamee

Figure 8-2). This evidence suggests the possibility that uncontrolled landscape fires burned through portions of the pinyon-juniper canopy, catastrophically de-stroyed at least one settlement, and destabilized the Red Hawk Wash watershed, which resulted in entrenchment of the channel. This interpretation is not unequivocal, however. For example, the expansion of wild-seed production areas into shrubby patches may have contributed more charcoal as higher population levels required increased food production after a.d. 1000. Yet another land use–based alternative is that “slash” (e.g., branches, bark) that accumulated from the processing of trees for architectural construction may have added to the mix of fuels combusted in landscape fires, thereby producing an increase in woody-fuel burning and charcoal deposition. Neither of these al-ternatives is incompatible with the crown-fire explanation and this highlights the potential vulnerability of these anthropogenic environments to devastating fires when climate and fuel conditions became optimal (Baker and Shinneman 2004; Floyd et al. 2004). Once alluvial deposition resumed on Simkins Flat prior to a.d. 1400, charcoal abundance increased, which indicates that fires may have remained somewhat frequent but became spatially discontinuous, creating a mosaic of trees, shrubs, and herbs attributable to longer fire-return intervals for the individual patches. In other words, the watershed’s fire regime was responding more to “natural” rather than anthropogenic factors. Increasing frequencies of fire-conducive cli-mate conditions may have driven short-lived increases in herbaceous pollen and soil phosphorus, thereby contributing to several prominent charcoal peaks prior to the second episode of erosion that occurred after a.d. 1650. The near-surface samples (above 20 cm depth) indicate a radically different environment of re-duced C4 plant inputs to soil organic matter (SOM), decreased herbaceous pollen and charcoal, and an increase in soil phosphorus. The elevated soil phosphorus is unlikely to be ash derived in light of the other proxy data. Rather, it may be attributable to artificial phosphorus enrichment from cattle grazing on the fan (Macphail et al. 2004). For the Forestdale Valley (Figures 8-4 and 8-5), the data indicate at least three different fire regimes between a.d. 1100 and a.d. 1910. Prior to a.d. 1400, large amounts of woody biomass were frequently burned (abundant charcoal, moder-ate to high soil phosphorus), probably during the spring or early summer, which increased sediment yields and produced fining-upward depositional units. Be-tween a.d. 1400 and a.d. 1600, fire frequency (decrease in soil phosphorus) and fire size (decrease in charcoal peak size) appear to have diminished in comparison to the period prior to a.d. 1400. Neither fire regime seems to have resulted in the expansion of herbaceous plants (pollen assemblages). After a.d. 1600, charcoal abundance and variability declined in the context of uniformly sandy sediment, increasing herbaceous pollen influx, higher soil phosphorus concentrations, and slightly greater input of C4 plants to SOM. These trends indicate a significant shift to very high–frequency fires for a given patch, with the possibility of a change in seasonality as well. One mechanism that could explain the change in sedimenta-tion to uniformly sandy alluvium would be an increase in rainfall intercept by groundcover (vegetation) during the monsoon, which would have reduced over-

Page 16: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

Systematic Indigenous Burning in the Upland Southwest 157

Figure 8-4. Stratigraphic data for locality 6 in the Forestdale Valley. Ap-proximate occupation periods are marked on the right (adapted from Roos 2008). Vertical dashed lines indicate overall mean values for grain size, char-coal, soil P, and δ13C.

FPO

Figure 8-5. Stratigraphic data for locality 10 in the Forestdale Valley. Ap-proximate occupation periods are marked on the right (adapted from Roos 2008). Vertical dashed lines indicate overall mean values for grain size, char-coal, soil P, and δ13C.

FPO

Page 17: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

158 C. I. Roos, A. P. Sullivan III, and C. McNamee

all sediment load. An increase in dormant-season burning of understory plants, a pattern consistent with fire-scar studies of Apache burning (e.g., Seklecki et al. 1996), would have increased herbaceous groundcover the following spring, thereby resulting in increased intercept of rainfall by vegetation. The minor in-crease in C4 plant contribution (stable carbon isotopes) relative to the increases in herbaceous pollen may indicate fall burning as well (Roos 2008).

Discussion

We may now evaluate the hypothesis presented above: was fire-re-gime variability attributable to climate exclusively? For pinyon-juniper wood-lands near the Grand Canyon, the answer is an emphatic “no.” The period of greatest predicted fire frequency in the climate reconstruction is after a.d. 1650 (see Figure 8-2). In contrast, the period of greatest fire frequency in the Simkins Flat stratigraphic record was between a.d. 800 and a.d. 1100/1200, although the climate reconstruction predicts average fire frequencies during this period. Pol-len-based climate interpretations for this period (a.d. 900–1300) would predict elevated Pinus pollen inputs because of hypothesized increase in the regularity of monsoon moisture (Davis and Shafer 1992; Peterson 1994). The Simkins Flat record documents an inverse pattern that we think is attributable to intensive hu-man promotion of wild, herbaceous vegetation. The charcoal, phosphorus, and carbon isotope data all corroborate the pollen data and suggest that this enriched understory vegetation was encouraged by the systematic application of fire to understory fuels. Furthermore, regular inputs of sedimentary charcoal after a.d. 1100 seem to indicate that the woodlands around Simkins Flat experienced a re-turn to “natural” fire regimes, which may have included patchy surface fires (cf. Baker and Shinneman 2004; Floyd et al. 2003). The pollen assemblage data imply that fires were not frequent or widespread enough to have a homogenizing effect on understory plants (e.g., conversion to grassy or herbaceous vegetation), but the charcoal evidence suggests that fires did occur regularly even after prehis-toric occupation terminated. For the ponderosa pine forests surrounding the Forestdale Valley, the climate hypothesis may be rejected as well. The three fire regimes inferred from the strati-graphic record are inconsistent with the predictions of the fire-climate reconstruc-tion. However, anthropogenic burning likely complemented natural fire regimes. Pre-planting burning of understory vegetation during the late prehistoric occupa-tion (prior to a.d. 1400) would have mimicked the timing and fuels of natural, pre-monsoon fires (Sullivan 1982). Increased agricultural burning would have accelerat-ed charcoal deposition (an increase in biomass combusted and possibly an increase in fire frequency) without necessarily promoting a herbaceous understory because domesticated plants would have occupied the newly burned patch. We should also note that there is no evidence for canopy reduction, as is predicted by the swidden agricultural burning model (Kohler 2004; Kohler and Mathews 1988). The dramatic shift in seasonality and ecological impact of fires before a.d. 1600 is not explicable in climatic terms either. However, the timing of this shift

Page 18: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

Systematic Indigenous Burning in the Upland Southwest 159

coincides with conventional interpretations of Western Apache colonization of the area (Whittlesey et al. 1997:198). As mentioned above, the Western Apache are known to have used fire in agriculture, hunting, and wild-plant management (Buskirk 1986; Gifford 1940). The shift to high-frequency autumn burning before a.d. 1600 is particularly consistent with postharvest burning of wild seed patches (Buskirk 1986:165–166), which may also have served to improve the predictabil-ity of hunting locations the following spring (Stewart 2002:221). An increase in burning during the dormant season (autumn, winter, or early spring) has been observed in fire-scar records of Apache burning elsewhere in the Southwest (Kaye and Swetnam 1999; Seklecki et al. 1996). These results are valuable for renewing interest in studies of subsistence in the prehistoric and protohistoric U.S. Southwest, which have tended to overem-phasize the importance of maize agriculture and its antiquity (e.g., Damp et al. 2002). Many of these studies give the impression that maize cultivation was the primary driver of subsistence-related activities of Southwestern peoples (Matson 1991; Matson and Chisholm 1991). In the Grand Canyon, archaeobotanical and pollen evidence from well-preserved “Pompeii-like” assemblages and geoarchae-ological evidence from Simkins Flat, which postdate the introduction of maize by millennia, do not support this assertion (Sullivan 1987; Sullivan and Ruter 2006). Additional studies of the diversity of Southwestern subsistence strategies will help us understand prehistoric decision making with respect to resource ranking (Barlow 2002), investigate opportunities and constraints in the “niche construc-tion” of prehistoric farmers and gatherer-gardeners (Rocek 1995), and advance the study of human impacts on past environments. These considerations are particularly important in the archaeological inves-tigation of resilience and sustainability (Redman 2005; van der Leeuw and Red-man 2002). In applied historical ecology studies today, recent (a.d. 1700–1900) fire-regime reconstructions have been used to argue for the reintroduction of fire, or the discontinuance of fire suppression, to Southwestern forests (e.g., Fulé et al. 1997). This chapter demonstrates that “natural,” lightning-driven ignitions and climate-driven fuel loads may not be sufficient to produce “healthy,” parklike forests in all climate conditions. In the Forestdale Valley, Apache burning during a period of frequent fire-conducive climate conditions (see Figure 8-2) propa-gated herbaceous ponderosa parkland and produced distinctive paleoecological signatures compared to contemporaneous unoccupied areas (Roos 2008), despite frequent natural and anthropogenic fires throughout the records. In contrast, the prehistoric high-frequency anthropogenic burning of herba-ceous vegetation in pinyon-juniper woodlands may have created an environment that was vulnerable to climatic change. Long-term drought that coincided with a multiyear warm period between a.d. 1066 and a.d. 1075 (Figure 8-6) combined with exceptional regional, antecedent moisture patterns to create a situation in which coarse fuels may have carried a catastrophic fire through the landscape, which destroyed settlements and destabilized the watershed. After abandon-ment, the ecosystem entered a new adaptive cycle (sensu Holling and Gunder-son 2002), in which a mosaic of herbs, shrubs, and trees was maintained by an entirely different fire regime.

Page 19: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

160 C. I. Roos, A. P. Sullivan III, and C. McNamee

Conclusion

Previous attempts to consider the feasibility of widespread anthropo-genic burning in the upland coniferous ecosystems of the pre-Hispanic South-west produced unconvincing results because they did not incorporate the full range of factors that influence changes in fire regimes and their effects on fuels, geochemistry, and plant communities. By incorporating archaeological and geo-archaeological evidence, in contrast, we rejected the hypothesis of exclusively climate-driven, “natural” fire histories and sustained the idea that prehistoric gatherer-gardeners of the Grand Canyon region used fire to increase the produc-tivity of herbaceous wild-plant communities in pinyon-juniper woodlands (Sul-livan 1996). Similarly, evidence from the Forestdale Valley tends to support the burn-plot hypothesis for Mogollon agriculture (Sullivan 1982) and the burning of surface fuels by Western Apache people to increase the productivity of herba-ceous wild-plant communities in ponderosa pine forests. The long-term ecological consequences of these anthropogenic environ-ments were somewhat variable. In each case, anthropogenic burning maintained consistent productive environments for generations. However, after more than two centuries of anthropogenic burning, extraordinary interannual moisture conditions, which coincided with a multiyear drought and multiyear warm pe-riod from a.d. 1066 to a.d. 1075, produced coarse-fuel loads that may have car-ried a settlement-destroying, watershed-altering canopy fire in the Upper Basin. Occupation of the Upper Basin continued for at least 100 years after this event, but the ecological consequences were apparently long lasting. Patch structure and vegetation composition never returned to previous conditions.

Figure 8-6. Reconstructed precipitation (Dean and Robinson 1977), tem-perature (Salzer and Kipfmueller 2005), and climate-predicted regional fire activity (Roos and Swetnam 2009) for the Grand Canyon area between a.d. 1050 and a.d. 1100.

FPO

Page 20: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

Systematic Indigenous Burning in the Upland Southwest 161

Along the Mogollon Rim, approximately 200 years of agricultural burning pro-duced modest landscape impacts until abandonment, destabilization, and erosion of upland gardens and fields occurred in the late fourteenth century a.d. Western Apache burning beginning in the sixteenth century a.d., in contrast, transformed the upland environments but did not increase their sensitivity to long-term droughts or unusual fire-climate conditions. Unlike the ecological consequences of low-intensity burning in the Upper Basin’s pinyon-juniper woodland, therefore, the long record of anthropogenic burning in the Forestdale Valley may have made the ponderosa pine forest there less vulnerable to fires caused by long-term climate change (Roos 2008). These examples illustrate that an understanding of the processes by which an-thropogenic environments are initiated and maintained, and an appreciation for the ecological consequences of these environments, such as variation in their stabil-ity, sustainability, and resilience, should be established and evaluated with a wide range of paleoenvironmental records. Anthropogenic landscape changes are neither uniformly detrimental nor beneficial. Even when the processes are similar, as in the cases of prehistoric Grand Canyon and protohistoric Forestdale Valley burning to promote wild-plant foods, the long-term ecological consequences may be dissimi-lar. By adopting a multidisciplinary paleoecological framework, archaeologists can contribute to the evolving discourse regarding human behavior and climate change and provide resource managers with the appropriate historical ecological contexts to evaluate alternative land use strategies in the present and for the future.

Notes

1. Because of the overwhelming abundance of charcoal in our sediments, we restricted ourselves further to counting macroscopic charcoal greater than 250 μm in size. 2. The following taxa divisions were used for partitioning pollen types to trees, herbs, and shrubs. For Simkins Flat, Juniperus, Pinus, Pseudotsuga, and Quercus were grouped as trees; Artemisia, Ephedra, and Rosaceae were grouped as shrubs; Com-positae, Chenopodiaceae-Amaranthaceae, Eriogonum, Poaceae, and Euphorbia were grouped as herbs. For Forestdale Valley, Abies, Cupressaceae, Picea, Podocarpus, Pinus, Quercus, and Juglans were grouped as trees; Ceonothus, Cercocarpus, Cowania, Prunus, Ephedra, Justicia, Purshia, Artemisia, Rosaceae, Sarcobatus, and Salix were grouped as shrubs; Ambrosia, Liguliflorae, Compositae, Chenopodiaceae-Amaranthaceae, Eriogonum, Graminae, Boorhavia, Cleome, Euphorbia, Umbelliferae, Plantago lanceolata, Erodium cicutarium, Lilliaceae, Cruciferae, and Tidestroemia were grouped as herbs.

Acknowledgments

We would like to thank Rebecca Dean for the opportunity to partici-pate in the Visiting Scholar Conference in 2007. Earlier versions of the research presented herein benefited from comments by Vance Holliday, Gary Huckleber-ry, Barbara Mills, and Mike Schiffer.

Page 21: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

162 C. I. Roos, A. P. Sullivan III, and C. McNamee

Geoarchaeological research in the Forestdale Valley was supported by an International Arid Lands Consortium grant to Vance Holliday (#05R-09). Institu-tional support for the AMS radiocarbon dating of detrital charcoal from the For-estdale Valley was provided by the NSF-supported Accelerator Mass Spectrom-etry Lab at the University of Arizona, Tim Jull, and Greg Hodgins. Jeff Homburg and Statistical Research Inc. generously provided access to lab space, chemicals, and the spectrophotometer for all soil phosphorus analyses. Funding for stable carbon isotope analysis was provided by a research grant from the Riecker Fund to Chris Roos. Research on the Fort Apache Indian Reservation would not have been possible without the support of the White Mountain Apache Tribal Council and Heritage Program. John Welch, Doreen Gatewood, Mark Altaha, and Nick Laluk were instrumental in providing much of this support. The Upper Basin Archaeological Research Project has been supported by grants from the University Research Council (University of Cincinnati), C. P. Taft Research Center and Taft Memorial Fund (University of Cincinnati), the McMicken College of Arts and Sciences (University of Cincinnati), the Center for Field Research (Earthwatch), and the USDA Forest Service (Kaibab National Forest). Alan Sullivan is most appreciative for the long-term intellectual and administrative support of the Upper Basin Archaeological Research Project provided by Dr. John A. Hanson, recently retired Kaibab National Forest ar-chaeologist. Calla McNamee’s geoarchaeological investigations at Simkins Flat were supported by the Geological Society of America and a University of Calgary Graduate Research grant.

References

Adams, Karen R.2004 Anthropogenic Ecology of the North American Southwest. In People and Plants

in Ancient Western North America, edited by Paul E. Minnis, pp. 167–204. Smithso-nian Books, Washington, D.C.

Allen, Craig D.2002 Lots of Lightning and Plenty of People: An Ecological History of Fire in the Up-

land Southwest. In Fire, Native Peoples, and the Natural Landscape, edited by Thomas R. Vale, pp. 143–193. Island Press, Washington, D.C.

2004 Ecological Patterns and Environmental Change in the Bandelier Landscape. In Archaeology of Bandelier National Monument: Village Formation on the Pajarito Plateau, New Mexico, edited by Timothy A. Kohler, pp. 19–67. University of New Mexico Press, Albuquerque.

Arno, Stephen F., and Kathy M. Sneck1977 A Method for Determining Fire History in Coniferous Forests of the Mountain West.

General Technical Report INT-42. USDA Forest Service Intermountain Research Station, Ogden, Utah.

Baker, William L., and Douglas J. Shinneman2004 Fire and Restoration of Piñon-Juniper Woodlands in the Western United States:

A Review. Forest Ecology and Management 189:1–21.

Page 22: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

Systematic Indigenous Burning in the Upland Southwest 163

Balling, Robert C., Jr., and Stephen G. Wells1990 Historical Rainfall Patterns and Arroyo Activity Within the Zuni River Drainage

Basin, New Mexico. Annals of the Association of American Geographers 80:603–617.Barlow, K. Renee

2002 Predicting Maize Agriculture Among the Fremont: An Economic Comparison of Farming and Foraging in the American Southwest. American Antiquity 67:65–88.

Barney, Milo A., and Neil C. Frischknecht1974 Vegetation Changes Following Fire in the Pinyon-Juniper Type of West-Central

Utah. Journal of Range Management 27:91–96.Biedenbender, Sharon H., Mitchel P. McClaren, Jay Quade, and Mark A. Weltz

2004 Landscape Patterns of Vegetation Change Indicated by Soil Carbon Isotope Composition. Geoderma 119:69–83.

Blackford, J. J.2000 Charcoal Fragments in Surface Samples Following a Fire and the Implications

for Interpretation of Subfossil Charcoal Data. Palaeogeography, Palaeoclimatology, Pa-laeoecology 164:33–42.

Bonnicksen, Thomas M.2000 America’s Ancient Forests: From the Ice Age to the Age of Discovery. John Wiley and

Sons, New York.Bonnicksen, Thomas M., M. Kat Anderson, Henry T. Lewis, Charles E. Kay, and Ruthann Knudson

1999 Native American Influences on the Development of Forest Ecosystems. In Eco-logical Stewardship: A Common Reference for Ecosystem Management, Vol. 2, edited by Nels C. Johnson, Andrew J. Malk, Robert Szaro, and William T. Sexton, pp. 439–470. Elsevier Science, Oxford, U.K.

Bowman, David M. J. S., G. D. Cook, and U. Zoppi2004 Holocene Boundary Dynamics of a Northern Australian Monsoon Rainforest

Patch Inferred from Isotopic Analysis of Carbon (14C and Delta13C) and Nitrogen (Delta15N) in Soil Organic Matter. Austral Ecology 29:605–612.

Brewer, D. G., R. K. Jorgensen, L. P. Munk, W. A. Robbie, and J. L. Travis1991 Terrestrial Ecosystem Survey of the Kaibab National Forest. USDA Forest Service,

Southwestern Region, Albuquerque, New Mexico.Brown, Peter M., Margot W. Kaye, Laurie S. Huckaby, and Christopher H. Baisan

2001 Fire History Along Environmental Gradients in the Sacramento Mountains, New Mexico: Influences of Local Patterns and Regional Processes. Ecoscience 8:115–126.

Buck, Caitlin E., J. Andrés Christen, and Gary N. James1999 Bcal: An Online Bayesian Radiocarbon Calibration Tool. Internet Archaeology

7. Electronic document, http://intarch.ac.uk/journal/issue7/buck/, accessed Sep-tember 1, 2009.

Buck, Caitlin E., J. B. Kenworthy, C. D. Litton, and A. F. M. Smith1991 Combining Archaeological and Radiocarbon Information: A Bayesian Ap-

proach to Calibration. Antiquity 65:808–821.Bull, William B.

1997 Discontinuous Ephemeral Streams. Geomorphology 19:227–276.Burroughs, William J.

2005 Climate Change in Prehistory: The End of the Reign of Chaos. Cambridge Univer-sity Press, Cambridge.

Buskirk, Winfred1986 The Western Apache: Living with the Land Before 1950. University of Oklahoma

Press, Norman.

Page 23: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

164 C. I. Roos, A. P. Sullivan III, and C. McNamee

Clark, James S.1988 Stratigraphic Charcoal Analysis on Petrographic Thin Sections: Application to

Fire History in Northwestern Minnesota. Quaternary Research 30:81–91.Clark, James S., Jason Lynch, Brian J. Stocks, and Johann Goldammer

1998 Relationships Between Charcoal Particles in Air and Sediments in West-Cen-tral Siberia. The Holocene 8:19–29.

Clark, James S., and P. Daniel Royall1995 Particle-Size Evidence for Source Areas of Charcoal Accumulation in Late Ho-

locene Sediments of Eastern North American Lakes. Quaternary Research 43:80–89.Connor, Edward F., and Daniel Simberloff

1986 Competition, Scientific Method, and Null Models in Ecology. American Scien-tist 74:155–162.

Cooper, Charles F.1960 Changes in Vegetation, Structure, and Growth of Southwestern Pine Forests

Since White Settlement. Ecological Monographs 30(2):130–164.Covington, W. Wallace

2003 The Evolutionary and Historical Context. In Ecological Restoration of Southwest-ern Ponderosa Pine Forests, edited by Peter Friederici, pp. 26–47. Island Press, Wash-ington, D.C.

Covington, W. Wallace, and Leonard F. DeBano1990 Effects of Fire on Pinyon-Juniper Soils. In Effects of Fire Management of South-

western Natural Resources: Proceedings of the Symposium, November 15–17, 1988, Tucson, Arizona, edited by Jay S. Krammes, pp. 78–86. General Technical Report RM-191. USDA Forest Service, Rocky Mountain Research Station, Fort Collins, Colorado.

Covington, W. Wallace, and Stephen S. Sackett1990 Fire Effects on Ponderosa Pine Soils and Their Management Implications. In

Effects of Fire Management of Southwestern Natural Resources: Proceedings of the Sympo-sium, November 15–17, 1988, Tucson, Arizona, edited by Jay S. Krammes, pp. 105–110. General Technical Report RM-191. USDA Forest Service, Rocky Mountain Research Station, Fort Collins, Colorado.

Crimmins, Michael A., and Andrew C. Comrie2004 Interactions Between Antecedent Climate and Wildfire Variability Across

South-Eastern Arizona. International Journal of Wildland Fire 13:455–466.Damp, Jonathan E., Stephen A. Hall, and Susan J. Smith

2002 Early Irrigation on the Colorado Plateau near Zuni Pueblo, New Mexico. American Antiquity 67:665–676.

Davis, Owen K., and David S. Shafer1992 A Holocene Climatic Record for the Sonoran Desert from Pollen Analysis of

Montezuma Well, Arizona, USA. Palaeogeography, Palaeoclimatology, Palaeoecology 92:107–119.

Dean, Jeffrey S.2004 Anthropogenic Environmental Change in the Southwest as Viewed from the

Colorado Plateau. In The Archaeology of Global Change: The Impact of Humans on Their Environment, edited by Charles L. Redman, Steven R. James, Paul R. Fish, and J. Daniel Rogers, pp. 191–207. Smithsonian Books, Washington, D.C.

Dean, Jeffrey S., Robert C. Euler, George J. Gumerman, Fred Plog, Richard H. Hevly, and Thor N. V. Karlstrom

1985 Human Behavior, Demography, and Paleoenvironment on the Colorado Pla-teaus. American Antiquity 50:537–554.

Page 24: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

Systematic Indigenous Burning in the Upland Southwest 165

Dean, Jeffrey S., and William J. Robinson1977 Dendroclimatic Variability in the American Southwest, A.D. 689 to 1970. U.S. Depart-

ment of Commerce Technical Information Service, PB-266 340. Springfield, Virginia.DeBano, Leonard F., Daniel G. Neary, and Peter F. Ffolliott

1998 Fire’s Effects on Ecosystems. John Wiley and Sons, New York.Denevan, William M.

1992 The Pristine Myth: The Landscape of the Americas in 1492. Annals of the As-sociation of American Geographers 82:369–385.

Dieterich, John H., and Thomas W. Swetnam1984 Dendrochronology of a Fire Scarred Ponderosa Pine. Forest Science 30:238–247.

Ezzo, Joseph A.1992 Dietary Change and Variability at Grasshopper Pueblo, Arizona. Journal of An-

thropological Archaeology 11:219–289.Floyd, M. Lisa, David D. Hanna, and William H. Romme

2004 Historical and Recent Fire Regimes in Piñon-Juniper Woodlands on Mesa Verde, Colorado, USA. Forest Ecology and Management 198:269–289.

Floyd, M. Lisa, William H. Romme, and David D. Hanna2003 Fire History. In Ancient Piñon-Juniper Woodlands: A Natural History of the Mesa

Verde Country, edited by M. Lisa Floyd, pp. 261–277. University of Colorado Press, Boulder.

Floyd, M. Lisa, William H. Romme, David D. Hanna, Mark Winterowd, Dustin Hanna, and John Spence

2008 Fire History of Piñon-Juniper Woodlands on Navajo Point, Glen Canyon Na-tional Recreation Area. Natural Areas Journal 28:26–36.

Fulé, Peter Z., W. Wallace Covington, and Margaret M. Moore1997 Determining Reference Conditions for Ecosystem Management of Southwest-

ern Ponderosa Pine Forests. Ecological Applications 7:895–908.Geib, Phil R., and Bruce B. Huckell

1994 Evidence for Late Preceramic Agriculture at Cibecue, East-Central Arizona. Kiva 59:433–454.

Gifford, Edward Winslow1940 Apache-Pueblo. Culture Element Distributions XII. University of California

Press, Berkeley.Gilman, Patricia A.

1997 Wandering Villagers: Pit Structures, Mobility, and Agriculture in Southeastern Ari-zona. Anthropological Research Paper No. 49. Department of Anthropology, Ari-zona State University, Tempe.

Grissino-Mayer, Henri D.1996 A 2129-Year Reconstruction of Precipitation for Northwestern New Mexico. In

Tree Rings, Environment, and Humanity, edited by Jeffrey S. Dean, David M. Meko, and Thomas W. Swetnam, pp. 191–204. Radiocarbon, special issue.

Grissino-Mayer, Henri D., William H. Romme, M. Lisa Floyd, and David D. Hanna2004 Climatic and Human Influences on Fire Regimes of the Southern San Juan

Mountains, Colorado, USA. Ecology 85:1708–1724.Grissino-Mayer, Henri D., and Thomas W. Swetnam

2000 Century-Scale Climate Forcing of Fire Regimes in the American Southwest. The Holocene 10:213–220.

Haury, Emil W.1985 Mogollon Culture in the Forestdale Valley, East-Central Arizona. University of Ari-

zona Press, Tucson.

Page 25: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

166 C. I. Roos, A. P. Sullivan III, and C. McNamee

1985 [1940] Excavations in the Forestdale Valley, East-Central Arizona. In Mogollon Culture in the Forestdale Valley, East-Central Arizona, pp. 135–279. University of Ari-zona Press, Tucson.

Haury, Emil W., and E. B. Sayles1985 [1947] An Early Pit House Village of the Mogollon Culture, Forestdale Valley,

Arizona. In Mogollon Culture in the Forestdale Valley, East-Central Arizona, by Emil W. Haury, pp. 220–371. University of Arizona Press, Tucson.

Hendricks, David M.1985 Arizona Soils. University of Arizona College of Agriculture, Tucson.

Hereford, Richard2002 Valley-Fill Alluviation During the Little Ice Age (ca. A.D. 1400–1880), Paria

River Basin and Southern Colorado Plateau, United States. Geological Society of America Bulletin 114:1550–1563.

Herr, Sarah A.2001 Beyond Chaco: Great Kiva Communities on the Mogollon Rim Frontier. Anthropo-

logical Papers of the University of Arizona No. 66. University of Arizona Press, Tucson.

Holling, C. S., and Lance H. Gunderson2002 Resilience and Adaptive Cycles. In Panarchy: Understanding Transformations in

Human and Natural Systems, edited by Lance H. Gunderson and C. S. Holling, pp. 25–62. Island Press, Washington, D.C.

Hopkins, Ralph L., and Kelcy L. Thompson2003 Kaibab Formation. In Grand Canyon Geology, 2nd ed., edited by Stanley S. Beus

and Michael Morales, pp. 196–211. Oxford University Press, Oxford.Hough, Walter

1903 Archaeological Field Work in Northeastern Arizona: The Museum-Gates Ex-pedition of 1901. In Annual Report of the U.S. National Museum, 1901, pp. 287–358. Government Printing Office, Washington, D.C.

Huckell, Lisa W.1999 Paleoethnobotany. In Living on the Edge of the Rim: Excavations and Analysis of

the Silver Creek Archaeological Research Project 1993–1998, Vol. 2, edited by Barbara J. Mills, Sarah A. Herr, and Scott Van Keuren, pp. 459–504. Arizona State Museum Archaeological Series No. 192. University of Arizona, Tucson.

Huckleberry, Gary A., and Brian R. Billman1998 Floodwater Farming, Discontinuous Ephemeral Streams, and Puebloan Aban-

donment in Southwestern Colorado. American Antiquity 63:595–616.Kaldahl, Eric J., and Jeffrey S. Dean

1999 Climate, Vegetation, and Dendrochronology. In Living on the Edge of the Rim: Excavations and Analysis of the Silver Creek Archaeological Research Project 1993–1998, Vol. 1, edited by Barbara J. Mills, Sarah A. Herr, and Scott Van Keuren, pp. 11–29. Arizona State Museum Archaeological Series No. 192. University of Arizona, Tuc-son.

Kaldahl, Eric J., Scott Van Keuren, and Barbara J. Mills2004 Migration, Factionalism, and the Trajectories of Pueblo IV Period Clusters in the

Mogollon Rim Region. In The Protohistoric Pueblo World, A.D. 1275–1600, edited by E. Charles Adams and Andrew I. Duff, pp. 85–94. University of Arizona Press, Tucson.

Karlstrom, Thor N. V.1988 Alluvial Chronology and Hydrologic Change of Black Mesa and Nearby Re-

gions. In The Anasazi in a Changing Environment, edited by George J. Gumerman, pp. 45–91. Cambridge University Press, Cambridge.

Page 26: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

Systematic Indigenous Burning in the Upland Southwest 167

Kaye, Margot W., and Thomas W. Swetnam1999 An Assessment of Fire, Climate, and Apache History in the Sacramento Moun-

tains, New Mexico. Physical Geography 20:305–330.Kitzberger, Thomas, Thomas W. Swetnam, and Thomas T. Veblen

2001 Inter-Hemispheric Synchrony of Forest Fires and El Niño-Southern Oscilla-tion. Global Ecology & Biogeography 10:315–326.

Kohler, Timothy A.2004 Pre-Hispanic Human Impact on Upland North American Southwestern Envi-

ronments: Evolutionary Ecological Perspectives. In The Archaeology of Global Change: The Impact of Humans on Their Environment, edited by Charles L. Redman, Steven R. James, Paul R. Fish, and J. Daniel Rogers, pp. 224–242. Smithsonian Books, Wash-ington, D.C.

Kohler, Timothy A., and Meredith H. Mathews1988 Long-Term Anasazi Land Use and Forest Reduction: A Case Study from South-

west Colorado. American Antiquity 53:537–564.Kruse, William H., Gerald J. Gottfried, Duane A. Bennett, and Humberto Mata-Manqueros

1996 The Role of Fire in Madrean Encinal Oak and Pinyon-Juniper Woodland De-velopment. In Effects of Fire on Madrean Province Ecosystems: A Symposium Proceed-ings, March 11–15, 1996, Tucson, Arizona, edited by Peter F. Ffolliott, Leonard F. De-Bano, Malchus B. Baker Jr., Gerald J. Gottfried, Gilberto Solis-Garza, Carleton B. Edminster, Daniel G. Neary, Larry S. Allen, and R. H. Hamre, pp. 99–106. General Technical Report RM-GTR-289. USDA Forest Service, Rocky Mountain Research Station, Fort Collins, Colorado.

Lynch, Jason A., James S. Clark, and Brian J. Stocks2004 Charcoal Production, Dispersal, and Redeposition from the Fort Providence

Experimental Fire: Interpreting Fire Regimes from Charcoal Records in Boreal For-ests. Canadian Journal of Forest Research 34:1642–1656.

McNamee, Calla2003 A Geoarchaeological Investigation of Simkin’s Flat, Upper Basin, Northern

Arizona. Unpublished M.A. thesis, Department of Archaeology, University of Cal-gary, Calgary, Alberta, Canada.

Macphail, Richard I., G. M. Cruise, Michael J. Allen, Johan Linderholm,and Peter Reynolds

2004 Archaeological Soil and Pollen Analysis of Experimental Floor Deposits; with Special Reference to Butser Ancient Farm, Hampshire, UK. Journal of Archaeological Science 31:175–191.

Mann, Charles C.2005 1491: New Revelations of the Americas Before Columbus. Knopf, New York.

Marlon, Jennifer, Patrick J. Bartlein, and Cathy Whitlock2006 Fire-Fuel-Climate Linkages in the Northwestern USA During the Holocene.

The Holocene 16:1059–1071.Matson, R. G.

1991 The Origins of Southwestern Agriculture. University of Arizona Press, Tucson.Matson, R. G., and Brian Chisholm

1991 Basketmaker II Subsistence: Carbon Isotopes and Other Dietary Indicators from Cedar Mesa, Utah. American Antiquity 56:444–459.

Merkle, John1952 An Analysis of a Pinyon-Juniper Community at Grand Canyon, Arizona. Ecol-

ogy 33:375–384.

Page 27: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

168 C. I. Roos, A. P. Sullivan III, and C. McNamee

Mills, Barbara J.1998 Migration and Pueblo IV Community Reorganization in the Silver Creek Area,

East-Central Arizona. In Migration and Reorganization: The Pueblo IV Period in the American Southwest, edited by Katherine A. Spielmann, pp. 65–80. Anthropological Research Papers No. 51. Arizona State University, Tempe.

Mills, Barbara J., Sarah A. Herr, and Scott Van Keuren (editors)1999 Living on the Edge of the Rim: Excavations and Analysis of the Silver Creek Archaeo-

logical Research Project 1993–1998. 2 vols. Arizona State Museum Archaeological Se-ries No. 192. University of Arizona, Tucson.

Morehead, Barbara J.1996 A Place Called Grand Canyon. University of Arizona Press, Tucson.

Morgan, Penelope, Guillermo E. Defossé, and Norberto F. Rodríguez2003 Management Implications of Fire and Climate Changes in the Western Ameri-

cas. In Fire and Climatic Change in Temperate Ecosystems of the Western Americas, ed-ited by Thomas T. Veblen, William L. Baker, Gloria Montenegro, and Thomas W. Swetnam, pp. 413–440. Ecological Studies 160. Springer, New York.

Newcomb, Joanne M.1999 Silver Creek Settlement Patterns and Paleodemography. In Living on the Edge of

the Rim: Excavations and Analysis of the Silver Creek Archaeological Research Project 1993–1998, Vol. 1, edited by Barbara J. Mills, Sarah A. Herr, and Scott Van Keuren, pp. 31–52. Arizona State Museum Archaeological Series No. 192. University of Arizona, Tucson.

Nordt, Lee C.2001 Stable Carbon and Oxygen Isotopes in Soils. In Earth Sciences and Archaeology,

edited by Paul Goldberg, Vance T. Holliday, and C. Reid Ferring, pp. 419–448. Klu-wer Academic, New York.

Periman, Richard D., Connie Reid, Matthew K. Sweifel, Gary McVicker, and Dan Huff1999 Human Influences on the Development of North American Landscapes: Ap-

plication for Ecosystem Management. In Ecological Stewardship: A Common Reference for Ecosystem Management, Vol. 2, edited by Nels C. Johnson, Andrew J. Malk, Robert Szaro, and William T. Sexton, pp. 471–492. Elsevier Science, Oxford, U.K.

Peterson, Kenneth Lee1994 A Warm and Wet Little Climatic Optimum and a Cold and Dry Little Ice Age

in the Southern Rocky Mountains, U.S.A. Climatic Change 26:245–269.Pieper, Rex D., and Roger D. Wittie

1990 Fire Effects in Southwestern Chaparral and Pinyon-Juniper Vegetation. In Ef-fects of Fire Management of Southwestern Natural Resources: Proceedings of the Sympo-sium, November 15–17, 1988, Tucson, Arizona, edited by Jay S. Krammes, pp. 87–93. General Technical Report RM-191. USDA Forest Service, Rocky Mountain Research Station, Fort Collins, Colorado.

Poulos, Helen Mills2007 Top Down and Bottom Up Influences on Fire Regimes, Diversity, Vegetation Patterns

in the Chihuahuan Desert Borderlands. Ph.D. dissertation, Yale University, New Ha-ven, Connecticut. University Microfilms, Ann Arbor.

Pyne, Stephen J.2000 Where Have All the Fires Gone? Fire Management Today 60(3):4–6.

Redman, Charles L.2005 Resilience Theory in Archaeology. American Anthropologist 107:70–77.

Redman, Charles L., Steven R. James, Paul R. Fish, and J. Daniel Rogers (editors)2004 The Archaeology of Global Change: The Impact of Humans on Their Environment.

Smithsonian Books, Washington, D.C.

Page 28: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

Systematic Indigenous Burning in the Upland Southwest 169

Reid, J. Jefferson, and Stephanie M. Whittlesey1999 Grasshopper Pueblo: A Story of Archaeology and Ancient Life. University of Ari-

zona Press, Tucson.Reimer, Paula J., Mike G. L. Baillie, Edouard Bard, Alex Bayliss, J. Warren Beck, Chanda J. H. Bertrand, Paul G. Blackwell, Caitlin E. Buck, George S. Burr, Kirsten B. Cutler, Paul E. Damon, R. Lawrence Edwards, Richard G. Fairbanks, Michael Friedrich, Thomas P. Guilderson, Alan G. Hogg, Konrad A. Hughen, Bernd Kromer, Gerry McCormac, Sturt Manning, Christopher Bronk Ramsey, Ron W. Reimer, Sabine Remmele, John R. Southon, Minze Stuiver, Sahra Talamo, F. W. Taylor, Johannes van der Plicht, and Constanze E. Weyhenmeyer

2004 IntCal04 Terrestrial Radiocarbon Age Calibration, 0–26 cal kyr BP. Radiocarbon 46:1029–1058.

Rocek, Thomas R.1995 Sedentarization and Agricultural Dependence: Perspectives from the Pithouse-

to-Pueblo Transition in the American Southwest. American Antiquity 60:218–239.Romme, William H., Craig D. Allen, John D. Bailey, William L. Baker, Brandon T. Bestelmeyer, Peter M. Brown, Karen S. Eisenhart, Lisa Floyd-Hanna, David W. Huffman, Brian F. Jacobs, Richard F. Miller, Esteban H. Muldavin, Thomas W. Swetnam, Robin J. Tausch, and Peter J. Weisberg

2008 Historical and Modern Disturbance Regimes, Stand Structures, and Landscape Dy-namics in Piñon-Juniper Vegetation of the Western U.S. Colorado Forest Restoration Institute, Colorado State University, Fort Collins.

Roos, Christopher I.2008 Fire, Climate, and Social-Ecological Systems in the Ancient Southwest: Alluvial Geo-

archaeology and Applied Historical Ecology. Ph.D. dissertation, University of Arizona, Tucson. University Microfilms, Ann Arbor.

Roos, Christopher I., and Thomas W. Swetnam2009 A Comparison of Two Millennial-Length Reconstructions of Climate Predicted

Fire Activity for Ponderosa Pine Forests of the Southern Colorado Plateau Region, Southwest US. Manuscript on file, Laboratory of Tree-Ring Research, University of Arizona, Tucson.

Salzer, Matthew W., and Kurt F. Kipfmueller2005 Reconstructed Temperature and Precipitation on a Millennial Timescale from

Tree-Rings in the Southern Colorado Plateau, U.S.A. Climatic Change 70:465–487.Schwartz, Douglas W.

1990 On the Edge of Splendor: Exploring Grand Canyon’s Human Past. School of Ameri-can Research Press, Santa Fe, New Mexico.

Seklecki, Mariette T., Henri D. Grissino-Mayer, and Thomas W. Swetnam1996 Fire History and the Possible Role of Apache-Set Fires in the Chiricahua Moun-

tains of Southeastern Arizona. In Effects of Fire on Madrean Province Ecosystems: A Symposium Proceedings, March 11–15, 1996, Tucson, Arizona, edited by Peter F. Ffol-liott, Leonard F. DeBano, Malchus B. Baker Jr., Gerald J. Gottfried, Gilberto Solis-Garza, Carleton B. Edminster, Daniel G. Neary, Larry S. Allen, and R. H. Hamre, pp. 238–246. General Technical Report RM-GTR-289. USDA Forest Service, Rocky Mountain Research Station, Fort Collins, Colorado.

Stewart, Omer C.2002 Forgotten Fires: Native Americans and the Transient Wilderness. University of

Oklahoma Press, Norman.Sullivan, Alan P., III

1982 Mogollon Agrarian Ecology. The Kiva 48:1–15.

Page 29: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

170 C. I. Roos, A. P. Sullivan III, and C. McNamee

1986 Prehistory of the Upper Basin, Coconino County, Arizona. Arizona State Museum Archaeological Series No. 167. University of Arizona, Tucson.

1987 Seeds of Discontent: Implications of a “Pompeii” Archaeobotanical As-semblage for Grand Canyon Anasazi Subsistence Models. Journal of Ethnobiology 7(2):137–153.

1992 Pinyon Nuts and Other Wild Resources in Western Anasazi Subsistence Econ-omies. Research in Economic Anthropology 6:195–239.

1996 Risk, Anthropogenic Environments, and Western Anasazi Subsistence. In Evolv-ing Complexity and Environmental Risk in the Prehistoric Southwest, edited by Joseph A. Tainter and Bonnie Bagley Tainter, pp. 145–167. Addison-Wesley, New York.

Sullivan, Alan P., III, Robert A. Cook, Matthew P. Purtill, and Patrick M. Uphus2001 Economic and Land-Use Implications of Prehistoric Fire-Cracked-Rock Piles,

Northern Arizona. Journal of Field Archaeology 28:367–382.Sullivan, Alan P., III, Philip B. Mink II, and Patrick M. Uphus

2002 From John W. Powell to Robert C. Euler: Testing Models of Grand Canyon’s Prehistoric Puebloan Settlement. In Culture and Environment in the American South-west: Essays in Honor of Robert C. Euler, edited by David A. Phillips Jr., and John A. Ware, pp. 49–68. SWCA Anthropological Research Paper No. 8. SWCA Environ-mental Consultants Inc., Phoenix, Arizona.

2007 Archaeological Survey Design, Units of Observation, and the Characterization of Regional Variability. American Antiquity 72:322–333.

Sullivan, Alan P., III, and Anthony H. Ruter2006 The Effects of Environmental Fluctuation on Ancient Livelihoods: Implica-

tions of Paleoeconomic Evidence from the Upper Basin, Northern Arizona. In Envi-ronmental Change and Human Adaptation in the Ancient Southwest, edited by David E. Doyel and Jeffrey S. Dean, pp. 180–203. University of Utah Press, Salt Lake City.

Swetnam, Thomas W., and Christopher H. Baisan1996 Historical Fire Regime Patterns in the Southwestern United States Since AD

1700. In Fire Effects in Southwestern Forests: Proceedings of the Second La Mesa Fire Sym-posium, Los Alamos, New Mexico, March 29–31, 1994, edited by Craig D. Allen, pp. 11–32. General Technical Report RM-GTR-286. USDA Forest Service, Rocky Moun-tain Research Station, Fort Collins, Colorado.

2003 Tree-Ring Reconstructions of Fire and Climate History of the Sierra Nevada and Southwestern United States. In Fire and Climatic Change in Temperate Ecosystems of the Western Americas, edited by Thomas T. Veblen, William L. Baker, Gloria Montenegro, and Thomas W. Swetnam, pp. 158–195. Ecological Studies 160. Springer, New York.

Swetnam, Thomas W., and Julio L. Betancourt1990 Fire-Southern Oscillation Relations in the Southwestern United States. Science

249:1017–1021.1998 Mesoscale Disturbance and Ecological Response to Decadal Climatic Variabil-

ity in the American Southwest. Journal of Climate 11:3128–3147.Vale, Thomas R.

2002 The Pre-European Landscape of the United States: Pristine or Humanized? In Fire, Native Peoples, and the Natural Landscape, edited by Thomas R. Vale, pp. 1–39. Island Press, Washington, D.C.

Vale, Thomas R. (editor)2002 Fire, Native Peoples, and the Natural Landscape. Island Press, Washington, D.C.

van der Leeuw, Sander, and Charles L. Redman2002 Placing Archaeology at the Center of Socio-Natural Studies. American Antiq-

uity 67:597–605.

Page 30: 8. Paleoecological Evidence for Systematic Indigenous ... › people.smu.edu › dist › ... · of ancient societies are incompletely understood, the “anthropogenic fire” ...

Systematic Indigenous Burning in the Upland Southwest 171

Welch, John R.1996 The Archaeological Measures and Social Implications of Agricultural Commitment.

Ph.D. dissertation, University of Arizona, Tucson. University Microfilms, Ann Arbor.Whitlock, Cathy, and R. Scott Anderson

2003 Fire History Reconstructions Based on Sediment Records from Lakes and Wet-lands. In Fire and Climatic Change in Temperate Ecosystems of the Western Americas, edited by Thomas T. Veblen, William L. Baker, Gloria Montenegro, and Thomas W. Swetnam, pp. 3–31. Ecological Studies 160. Springer, New York.

Whitlock, Cathy, and Patrick J. Bartlein2004 Holocene Fire Activity as a Record of Past Environmental Change. In The Qua-

ternary Period in the United States, edited by A. R. Gillespie, S. C. Porter, and B. F. Atwater, pp. 479–490. Elsevier, New York.

Whitlock, Cathy, and Sarah H. Millspaugh1996 Testing the Assumptions of Fire-History Studies: An Examination of Modern

Charcoal Accumulation in Yellowstone National Park, USA. The Holocene 6:7–15.Whitlock, Cathy, Carl N. Skinner, Patrick J. Bartlein, Thomas Minckley, and Jerry A. Mohr

2004 Comparison of Charcoal and Tree-Ring Records of Recent Fires in the Eastern Klamath Mountains, California, USA. Canadian Journal of Forest Research 34:2110–2121.

Whittlesey, Stephanie M., William L. Deaver, and J. Jefferson Reid1997 Yavapai and Western Apache Archaeology of Central Arizona. In Vanishing Riv-

er, edited by Stephanie M. Whittlesey, pp. 185–214. Statistical Research, Inc., Tucson.Williams, Gerald W.

2002 Aboriginal Use of Fire: Are There Any “Natural” Plant Communities? In Wil-derness and Political Ecology: Aboriginal Influences and the Original State of Nature, ed-ited by Charles E. Kay and Randy T. Simmons, pp. 179–214. University of Utah Press, Salt Lake City.

Wills, W. H.1995 Archaic Foraging and the Beginning of Food Production in the American

Southwest. In Last Hunters-First Farmers: New Perspectives on the Transition to Agri-culture, edited by T. Douglass Price and Anne Birgitte Gebauer, pp. 215–242. School of American Research, Santa Fe, New Mexico.

Wittie, Roger D., and Kirk C. McDaniel1990 Effects of Tebuthiuron and Fire on Pinyon-Juniper Woodlands in Southcen-

tral New Mexico. In Effects of Fire Management of Southwestern Natural Resources: Proceedings of the Symposium, November 15–17, 1988, Tucson, Arizona, edited by Jay S. Krammes, pp. 174–179. General Technical Report RM-191. USDA Forest Service, Rocky Mountain Research Station, Fort Collins, Colorado.

Wyckoff, Don G.1977 Secondary Forest Succession Following Abandonment of Mesa Verde. The Kiva

42:215–231.