29
37 Introduction In many parts of North America, as other chapters in this volume demon- strate, the change from Paleoindian to Archaic lifeways is demonstrably a transformation taking place within a single cultural tradition. In the southern interior of the Pacific Northwest, however, this event is marked by two tradi- tions occurring in sequence. The Western Stemmed Tradition, which is the regional manifestation of Paleoindian, is followed by the Old Cordilleran Tra- dition, which marks the beginning of the Archaic. The two differ markedly in stone, bone and processing technologies, subsistence, tool and clothing styles, land use, settlement, and even the morphology of the participants themselves. In this paper we present the characteristics of the two traditions, explore their chronological distributions, and offer new information from Beech Creek, a site containing both traditions in an apparently continuous chronological se- quence. We then consider possible explanations for this behavioral and mor- phological disjunction, finding ethnic replacement to be a better explanation than in situ evolution for the entire suite of changes. Both alternatives, to some extent, look to Early Holocene climatic change as a primary driver. The Region and Its Changing Environment The area addressed in this paper extends along the Northwest Coast of North America from Glacier Bay in Alaska to the Oregon–California border, and includes the interior basins of the Columbia and Fraser Rivers (fig. 3.1). Our discussion also reaches into southeastern Oregon, an area usually consid- ered part of the Great Basin. Topographically, the region is dominated by the north–south trending Coast, Cascade, and Rocky Mountain Ranges, and their intervening valleys. Between the Coast and Cascade Ranges in Washington and Oregon lie the Puget Lowland and Willamette Valley. The Columbia and Fraser Plateaus fill the broad expanse between the Cascades and Rockies. Unlike most of the regions discussed in this volume, the Northwest is a land of marked topographic, climatic, and ecological variability. Climate is maritime; warm dry summers contrast with cool, wet winters, when moisture-laden clouds flow eastward off the Pacific Ocean and back up against the mountain ranges. The land west of the Cascade and Canadian Coast ranges, therefore, receives most of the precipitation, which supports dense, mixed-conifer for- ests. This closed canopy restricts terrestrial food resources to shorelines, re- covering burns, subalpine and alpine environments, and a few prairies; most of the resources this region had to offer foraging peoples were marine, in- cluding hyperabundant, anadromous salmonids. The rain shadow from the western mountains prevents much of the moisture from reaching the interior plateaus, which range from arid in south-central Oregon to subhumid in the northern Fraser Plateau. Conifer forests occupy the uplands, shrub steppe the lowlands. With their more open plant communities, the intermountain plateaus offer better access to terrestrial game, but salmon constituted much of the animal biomass of these drier lands. During the terminal Pleistocene, Cordilleran and montane glaciers covered all but the southern lowlands of Oregon, Washington, and Idaho. Deglaciation had begun by ca 17,000 cal B.P. (Waitt and Thorsen 1983) and had receded north of the Canadian border by 15,000 cal B.P. By 12,000 cal B.P., all but the highest mountains of Alberta and British Columbia were fully deglaciated The Paleoindian to Archaic Transition in the Pacific Northwest In Situ Development or Ethnic Replacement? James C. Chatters, Steven Hackenberger, Anna M. Prentiss, and Jayne-Leigh Thomas 3 TAM Bousman 13106 complete.indd 37 TAM Bousman 13106 complete.indd 37 4/30/12 8:26 AM 4/30/12 8:26 AM Not for reproduction or distribution Not for reproduction or distribution Not for reproduction or distribution Not for reproduction or distribution Not for reproduction or distribution

The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

37

Introduction

In many parts of North America, as other chapters in this volume demon-strate, the change from Paleo indian to Archaic lifeways is demonstrably a transformation taking place within a single cultural tradition. In the southern interior of the Pacifi c Northwest, however, this event is marked by two tradi-tions occurring in sequence. The Western Stemmed Tradition, which is the regional manifestation of Paleo indian, is followed by the Old Cordilleran Tra-dition, which marks the beginning of the Archaic. The two differ markedly in stone, bone and processing technologies, subsistence, tool and clothing styles, land use, settlement, and even the morphology of the participants themselves. In this paper we present the characteristics of the two traditions, explore their chronological distributions, and offer new information from Beech Creek, a site containing both traditions in an apparently continuous chronological se-quence. We then consider possible explanations for this behavioral and mor-phological disjunction, fi nding ethnic replacement to be a better explanation than in situ evolution for the entire suite of changes. Both alternatives, to some extent, look to Early Holocene climatic change as a primary driver.

The Region and Its Changing Environment

The area addressed in this paper extends along the Northwest Coast of North America from Glacier Bay in Alaska to the Oregon–California border, and includes the interior basins of the Columbia and Fraser Rivers (fi g. 3.1). Our discussion also reaches into southeastern Oregon, an area usually consid-ered part of the Great Basin. Topographically, the region is dominated by the north–south trending Coast, Cascade, and Rocky Mountain Ranges, and their intervening valleys. Between the Coast and Cascade Ranges in Washington and Oregon lie the Puget Lowland and Willamette Valley. The Columbia and Fraser Plateaus fi ll the broad expanse between the Cascades and Rockies.

Unlike most of the regions discussed in this volume, the Northwest is a land of marked topographic, climatic, and ecological variability. Climate is maritime; warm dry summers contrast with cool, wet winters, when moisture- laden clouds fl ow eastward off the Pacifi c Ocean and back up against the mountain ranges. The land west of the Cascade and Canadian Coast ranges, therefore, receives most of the precipitation, which supports dense, mixed- conifer for-ests. This closed canopy restricts terrestrial food resources to shorelines, re-covering burns, subalpine and alpine environments, and a few prairies; most of the resources this region had to offer foraging peoples were marine, in-cluding hyperabundant, anadromous salmonids. The rain shadow from the western mountains prevents much of the moisture from reaching the interior plateaus, which range from arid in south- central Oregon to subhumid in the northern Fraser Plateau. Conifer forests occupy the uplands, shrub steppe the lowlands. With their more open plant communities, the intermountain plateaus offer better access to terrestrial game, but salmon constituted much of the animal biomass of these drier lands.

During the terminal Pleistocene, Cordilleran and montane glaciers covered all but the southern lowlands of Oregon, Washington, and Idaho. Deglaciation had begun by ca 17,000 cal B.P. (Waitt and Thorsen 1983) and had receded north of the Canadian border by 15,000 cal B.P. By 12,000 cal B.P., all but the highest mountains of Alberta and British Columbia were fully deglaciated

The Paleo indian to Archaic Transition in the Pacifi c NorthwestIn Situ Development or Ethnic Replacement?James C. Chatters, Steven Hackenberger, Anna M. Prentiss, and Jayne- Leigh Thomas 3

TAM Bousman 13106 complete.indd 37TAM Bousman 13106 complete.indd 37 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 2: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

38 Chatters, Hackenberger, Prentiss, and Thomas

(Clague 1981). Sea levels rose rapidly between 17,000 and 9000 cal B.P. due to a combination of isostatic depression and increased ocean volume (Mann and Hamilton 1995; Fedje et al. 1996). Along the north coast, this led to trans-gression above modern sea levels, increasing the convolution of the coast, and potentially meaning greater littoral productivity (Moss and Erlandson 1996). Westerly air fl ow from wasting Laurentide ice, however, reduced upwelling along the southern coast until after ca 9000 cal B.P. (Sancetta et al. 1992). This, along with constantly changing shorelines, meant lower nearshore ma-rine productivity. Although salmon were well established in some parts of the Fraser and Columbia systems by 9000 cal B.P. (Butler and Connor 2004; Matson 1996), they would not be abundant in inland rivers until thousands of years later (Chatters et a1. 1995, Butler and Chatters 2003).

In the late Pleistocene and for the fi rst three millennia of deglaciation, a steppe- tundra similar to the Beringian mammoth steppe (Guthrie 1982) occu-pied exposed land below the ice margins, except in southern Oregon, where conifer woodland and grasslands existed around pluvial lakes. As tempera-

Figure 3.1. Map of the Pacifi c Northwest of North America, showing

the areas and key sites discussed in this chapter.

TAM Bousman 13106 complete.indd 38TAM Bousman 13106 complete.indd 38 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 3: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

39 Pacifi c Northwesttures increased, a lodgepole pine, spruce, and mountain hemlock woodland akin to Canadian taiga developed in the Puget and Willamette lowlands, to be replaced by open oak and Douglas fi r parklands after 10,000 cal B.P. (Barnosky et al. 1987). Denser conifer woodlands and forests could not be found in the southern Cascades below 1000 meters on the west fl ank and 1500 meters on the east (Chatters 1998a). Dry spruce and pine parklands then occupied the northern Cascades (Heinrichs et al. 2002; Pellatt et al. 1998). Upper timber lines throughout the region stood as much as 300 meters above modern levels (Clague and Mathewes 1989). After 9500 cal B.P., montane forests became more mesic and moved downslope (Barnosky et al. 1987; Hebda 1995), but the Puget Lowland and Willamette Valley remained open or sparsely forested until after 7000 cal B.P. In the intermountain plateaus, conditions became increasingly arid as the Early Holocene progressed (Chatters 1998b). The ar-eas now occupied by shrub steppe supported grassland from 12,000 cal B.P. until between 9500 and 9000 cal B.P., when an arid sagebrush and chenopod steppe became widespread. In the Rockies, true forests were rare throughout the Early Holocene, except in the higher Canadian portions of the range. Much of what is today pine, fi r, or cedar–hemlock forest was either unfor-ested or covered in steppe–forest mosaic (Chatters 1996). Animals that inhab-ited these plant communities included now- extinct megafauna until between 14,000 and 13,000 cal B.P. (Lyman and Livingston 1983), but, except for Bison antiquus (Irwin and Moody 1978), only modern fauna survived thereafter (Chatters 1998b). The largest modern herbivores—elk and bison—were com-mon in the Columbia River Basin before 9000 cal B.P., but declined sharply thereafter (Gustafson 1972).

Patterns of vegetation change and weakened nearshore marine upwell-ing indicate that the climate before 9000 cal B.P. was more continental, with marked seasonality and a more easterly air fl ow. Higher timber lines are evi-dence of hotter summers; high frequencies of rockfall in caves and rockshel-ters on the Columbia Plateau result from colder winters. Patterns of vegeta-tion distribution—the change from xeric to mesic parklands west of the Coast and Cascade ranges, downslope movement of forests in the mountains them-selves, a change from grasslands to shrub steppe on the plateaus, and contin-ued high tree lines—show continued warmth but a shift to a more maritime climatic pattern after 9000 cal B.P. (Chatters 1998b).

The change from continental to maritime climate altered the resource potentials on either side of the coastal mountain ranges. Parklands west of the mountains meant higher game and food plant productivity before 9000 cal B.P., at a time marine productivity was reduced along the southern coast. After that time, terrestrial productivity remained high while marine produc-tivity was on the increase. East of the mountains, expanding forests and the loss of grasslands meant a decline in terrestrial productivity, making riverine resources proportionately more important. Change to maritime climate meant promotion of some food species, particularly geophytes, which thrive on winter moisture. These changes had important implications for early hunter- gatherers.

Paleo indian and Archaic Lifeways

The archaeological record of the Northwest begins as early as 14,000 cal B.P., if the speared mastodon found at Manis in northwest Washington (Gustafson et al. 1979) and human coprolites from Paisley Caves in south- central Oregon (Gilbert et al. 2008) are taken at face value. A light scattering of isolated Clovis and other fl uted projectile points has been reported south of the Canadian border (Carlson 1996a; Moss and Erlandson 1996), but the Richie- Roberts Cache in central Washington (Mehringer 1988) and the Dietz Site in Oregon (Fagan 1988) are the only discrete assemblages of this early material. The earliest well- represented cultural traditions in the Northwest are the Western Stemmed and Old Cordilleran traditions, which were both well established by 12,000 cal B.P. What we know in detail about the behavior of people who left behind the artifacts of these traditions comes primarily from the southern interior of our region, where site records have not been diminished by the

TAM Bousman 13106 complete.indd 39TAM Bousman 13106 complete.indd 39 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 4: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

40 Chatters, Hackenberger, Prentiss, and Thomas

acid soils and extensive bioturbation of conifer forests. In that subregion, the two traditions occur in sequence: Western Stemmed Tradition (WST) is our Paleo indian; the Old Cordilleran Tradition (OCT) ushers in the Archaic. They differ in their tool, food processing, and clothing technologies, land use and settlement strategies, and subsistence emphases.

Western Stemmed Tradition

Manifestations of the WST are found from Banff in the Northern Rockies to the northwestern coast of Oregon and extend southward into the Great Basin, where they have become known as the Paleo archaic (See Jones and Beck, this volume). They are known by such chronological labels as the Youngs River Complex on the Lower Columbia River (Minor 1984), Philipi Phase on the middle Columbia River (Dumond and Minor 1983), Windust Phase on the Columbia Plateau (Leonhardy and Rice 1970), the Goatfels Complex on the upper Kootenai (Choquette 1996), Banff II in southwestern Alberta (Fedje 1996), and Haskett in southern Idaho (Butler 1965). Components attribut-able to this tradition range from 13,550 cal B.P. at the Cooper’s Ferry Site in northern Idaho to as late as 9000 cal B.P. (table 3.1). Most date between 13,000 and 9500 cal B.P.

Technologies

Lithic and bone technologies characteristic of the WST are complex. Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points, biface cores that also become projectile or knife performs, large end and side scrapers produced on large fl akes with little edge modifi cation,

Table 3.1. Western Stemmed Tradition (Paleo indian) site chronology in the Northwest.

Site 14C Age Range Calibrated Age

cal B.P.a Reference Map No.

Banff: Vermillion and Eclipse sites

10,200 ± 609600 ± 60

12,028–11,82011,102–10,790

Fedje 1996 916

Lind Coulee 10,250 ± 409810 ± 40

12,083–11,83711,241–11,204

Craven 2003 8

Sentinel Gap 10,180 ± 60 11,999–11,758 Galm and Gough 2001, 2008 10Wildcat Canyon 10,600 ± 200

7370 ± 19012,832–12,240

8362–8016Cole 1965Dumond and Minor 1983

6

Marmes Rockshelter 10,810 ± 3008525 ± 100

13,094–12,3869604–9426

Sheppard et al. 1987Hicks 2004

3

Marmes Floodplain 10,130 ± 3009710 ± 40

12,341–11,24911,203–11,126

Sheppard et al. 1987Hicks 2004

3

Wewukiyepuh 10,270 ± 50 12,134–11,840 Sappington and Schuknecht- McDaniel 2001

25

Hatwai I 10,820 ± 1408830 ± 310

12,926–12,57010,249–9528

Ames et al. 1981 2

10NP453 8919 ± 60 10,174–9919 Ridenour 2006 25Cooper’s Ferry 11,370 ± 70

8410 ± 7013,300–13,179

9518–9321Davis and Sisson 1998 1

Tucker 9520 ± 609160 ± 50

11,069–10,70010,385–10,198

Pinson 2004 17

35LK1881 8940 ± 60 10,196–9932 Oetting 1994 2435LK2076 8810 ± 100 10,133–9688 Oetting 1994 26Cougar Mt. Cave 8510 ± 250 9887–9136 Ferguson and Libby 1962 29Ft. Rock Cave 10,200 ± 230

8550 ± 15012,346–11,231

9743–9318Bedwell 1973 11

Connely Cave #5B 9540 ± 260 11,204–10,514 Bedwell 1973 18Paulina Lake 9970 ± 470b 12,347–10,785 Connolly and Jenkins 1999b 13Dirty Shame Rockshelter 9500 ± 95

7850 ± 12010,071–10,606

8972–8523Aikens et al. 1977Hanes 1988

19

Harney Lake 8680 ± 55 9680–9551 Gehr 1980 28Hetrick 9850 ± 110 11,599–11,163 Rudolph 1995 14Buhl 10,675 ± 95 12,830–12,641 Green et al. 1998 4Shoup Rockshelter 8125 ± 230 9398–8727 Swanson and Snead 1966 38Redfi sh Overhang 10,100 ± 300

8060 ± 285 12,337–11,230

9290–8599 Sargeant 1973 12

a. One- sigma range, determined using Calib5.b. Pre- Mazama sediments were arbitrarily divided into components, so separation between Western Stemmed and Old Cordilleran components cannot be made. No upper age limit can be given.

TAM Bousman 13106 complete.indd 40TAM Bousman 13106 complete.indd 40 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 5: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

41 Pacifi c Northwest

keeled scrapers or scraper- planes, denticulate scrapers, burins, gravers, drills, and a wide array of retouched and utilized fl akes (fi g. 3.2). Cobble imple-ments, milling stones and manos occur, but infrequently. Two tools that are unique to the WST are chipped stone crescents and fi nely crafted, egg- sized, plumb bob–shaped stones with longitudinal grooves, known as bola stones. Projectile points, which were produced through bifacial percussion reduction and fi nished with broad, collateral fl akes, appear to have been made long and broad (e.g., Butler 1965; Rice 1972), but were typically resharpened to stemmed stubs while still in the haft. This intensity of reworking and a ten-dency to exhibit bending fractures at the stem–blade contact are thought to indicate use as knives (Galm and Gough 2008), a common characteristic of Paleo indian technology (e.g., Ahler 1971).

A complex bone technology is also found when conditions allow good preservation. Single- piece and composite bone shafts, unilaterally barbed har-poon points, atlatl spurs, wedges, and other implements have been found at multiple sites (Galm and Gough 2008; Irwin and Moody 1978; Leonhardy 1970; Rice 1972). Bone needles, little larger than modern sewing needles, occur in most early components of this tradition, another link to Paleo indian toolkits continent- wide. Harpoon points, crescents, bolas, atlatl spurs, and multi- piece bone shafts all represent compound tools, indicating a complex technology with a high degree of planning depth (Torrence 1983).

In contrast to the complex toolkit, food processing technology appears to have been quite limited. The presence of a small number of lightly utilized milling stones at such sites as Marmes Rockshelter (Rice 1969; Hicks 2004) and the Vine Site (Lewarch and Benson 1989), and a high frequency of such tools at the Goldendale Site (Warren et al. 1963) provide evidence of what is thought to have been seed processing, although no direct evidence of plant use has yet been obtained. Conspicuous in their absence or rarity are earth ovens and boiling stones. No oven feature has yet been found in association with, or contemporaneous with WST components. Thermally altered rock is rare and, when present, consists of large fragments consistent with cooking by roasting rather than boiling (e.g., Sappington and Schuknecht- McDaniel 2001).

One fi nal aspect of technology is the use of the body as a tool. Only two complete skulls are known from this tradition in the southern interior: Ken-newick Man, from eastern Washington (Chatters 2000), and the Buhl woman from southern Idaho (Green et al. 1998). In Kennewick Man, wear on the

Figure 3.2. Stone and bone implements characteristic of the Western Stemmed Tradition in the Pacifi c Northwest. Included are a large biface (a), end scraper (b) edge- modifi ed fl akes (c, d), keeled scraper or scraper plane (e), bola stone (f), projectile points (g–i), crescent (j), atlatl spur (k), needles (1), bead (m), bone foreshaft (n), and unilaterally barbed harpoon (o). Items a–e, h–j, and 1–o are from Lind Coulee (Irwin and Moody 1978); f is from Lake Pend Oreille (see Weisz 2006), g from the Kennewick Man Site, k from Granite Point (see Leonhardy 1970). Drawings f, g, k by Claire Chatters; all others courtesy Washington State University

TAM Bousman 13106 complete.indd 41TAM Bousman 13106 complete.indd 41 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 6: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

42 Chatters, Hackenberger, Prentiss, and Thomas

lingual surfaces of his lower incisors indicates the front teeth had a paramas-ticatory grasping function. Buhl, who died as a young woman, also had heav-ily worn anterior teeth (Green et al. 1998). High anterior dental wear is also found in other Paleo american individuals in the western United States (Chat-ters 2008). This use of the mouth may relate to the lack of food processing technology, a diet of at least partially dried or tough, partially cooked meat, and the evidence that projectiles doubled as knives. It is entirely possible that WST people held tough meat in their teeth to cut it, much as Eskimo peoples are known to have done.

Settlement Strategy

Although it was long believed that WST peoples concentrated their settle-ments along river corridors (Ames 1988, Ames et al. 1998), it has recently be-come clear that they made much more extensive use of the entire landscape. Sites like Rock Island Overlook (Valley 1975), Marmes Rockshelter (Rice 1969) and Five- Mile Rapids (Cressman et a1. 1960) have indeed been found on the region’s trunk streams, but sites also occur in upland areas adjacent to rivers, as at Sentinel Gap (Galm and Gough 2008) and the Portland (Oregon) Basin; at wetlands in the intermountain plateaus in such sites as Lind Cou-lee (Daugherty 1956), Willow Lakes, and Bishop Springs (Huckleberry et al 2003); and well into the Cascade and Rocky Mountains at sites like Judd Peak Rockshelter (Daugherty et al. 1987a), Vine (Lewarch and Benson 1989), and Pilcher Creek (Brauner 1985). Inland wetlands and lakes seem to have been particularly favored. WST isolates and sites have been found at Pend Oreille (Miss and Huson 1987), Palmer (Salo 1987), and Goose lakes. In southeastern Oregon, settlements near Early Holocene lakes are the rule (e.g., Pinson 2004; Connolly and Jenkins 1999a, 199b), which led to their inclusion in Bedwell’s (1973) Western Pluvial Lakes Tradition.

WST sites show a tendency for frequent reuse, as if they were nodes in an annual subsistence round. Lind Coulee was repeatedly used as a hunt-ing camp, from which people focused on bison and elk (Irwin and Moody 1978). The Goldendale Site was a seed gathering camp (Warren et al. 1963), probably used in fall; Five- Mile Rapids and Rock Island Overlook are located at prime locations for harvesting either fi sh or the birds that fed on their car-casses. Marmes Rockshelter was probably a winter encampment. These sites appear to have been base camps from which foraging forays were made in a logistical- like pattern. In his analysis of early Columbia Plateau assemblages, Ames (1988) found that WST sites could be grouped into those with many tool types and those with few, supporting the notion that both base camps and fi eld camps were utilized. One of the best examples of a fi eld camp is Wewu-kiyepuh, in northern Idaho (Sappington and Schuknecht- McDaniel 2001), which included remains of a single elk, a hearth, nine tools including two broken projectile points, and a scatter of debitage from producing replacement points out of imported bifaces.

On a regional scale, individual artifact types also show a high degree of patterning, indicative of systematic use of habitats. Crescents, for example, have only been found in upland settings, near wetlands. Although the only reported specimens come from Lind Coulee (Daugherty 1956), collectors have also found them at Willow Lakes and at upland spring sites (Lenz in press). Bolas, conversely, have been found almost exclusively in lowland settings—along rivers and their nearby terraces. Numerous examples have been found at the mouth of Pend Oreille Lake (Weisz 2006), Rock Island Rapids (Valley 1975), Five- Mile Rapids (Cressman et al. 1960), and as isolates on the ter-races of the Portland Basin (Pettigrew 1981). The only exception is the Pilcher Creek Site, high in the Blue Mountains of northeast Oregon, where the bola stones appear to have been manufactured (Brauner 1985).

Social groups were small and mobile. On the few occasions where single event camps are found—or can be discerned—as at Lind Coulee (Irwin and Moody 1978) and Sentinel Gap (Galm and Gough 2008), they cover at most a few hundred square meters. Feature and artifact patterning is discrete—evidence of short- term occupancy and, therefore, high residential mobility.

TAM Bousman 13106 complete.indd 42TAM Bousman 13106 complete.indd 42 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 7: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

43 Pacifi c NorthwestIn keeping with high mobility, WST people left few traces of dwellings (but see Connolly 1999a, 1999b). A comparison of the frequencies of expedient stone tools (e.g., utilized and retouched fl akes, hammerstones, cobble tools) to curated tools (e.g., projectile points, bifaces, scrapers, gravers, drills), how-ever, indicates that WST social groups in the interior northwest were less mobile than their OCT successors. We compared WST and OCT assemblages from Granite Point and Beech Creek, the two sites that allow clear separation of assemblages from the two traditions. In both cases, the WST assemblages were much more expedient,1 indicating a lower level of mobility (Hayden et al. 1996).

Subsistence

Although seed use is inferred from the presence of milling stones, only the an-imal part of the WST diet is known. WST faunal assemblages are highly vari-able; those in south- central Oregon tend to emphasize rabbits and birds (e.g., Oetting 1994; Pinson 2004), whereas those from more northern contexts tend to be dominated by larger game, while including a range of smaller prey. Deer, elk, and bison rank fi rst or second in NISP in nearly all northern sites with reported faunal assemblages, including Hetrick (Rudolph 1995), Hatwai (Atwell 1989), Wewukiyepuh (Sappington and Schuknecht- McDaniel 2001), Lind Coulee2 (Irwin and Moody 1978), Sentinel Gap (Gough pers. comm.), Granite Point, Marmes Rockshelter (Gustafson 1972), Windust Caves (Rice 1965), and Wildcat Canyon (Dumond and Minor 1983). Birds and rabbits or marmots often rank in the top four. Fish are not a signifi cant part of any WST assemblage on the Columbia Plateau.3 There is no sign of food storage, although it is likely that meat was at least partially dried to extend its use life.

Mortuary Practices

Few sets of human remains have been found in association with WST assem-blages in the northwest or elsewhere. Only four are known—from Marmes Rockshelter (Krantz 1979), Buhl (Green et al. 1998), and Kennewick (Chat-ters 2000), and the imprecisely provenienced Stick Man (Chatters et al. 2000). Little patterning is discernible from this group. The Marmes remains, from undoubted WST contexts, appear to have been secondary cremations (Chat-ters 2010a), deposited repeatedly in the same corner of the rockshelter over a long period of time. As such, they are yet another indicator of WST’s pat-terned use of the landscape beginning as early as 12,000 cal B.P. Buhl, also from before 12,000 cal B.P., was alone and appears to have been abandoned or secondarily deposited. Kennewick Man, whether interred (Huckleberry and Stein 1999) or abandoned (Chatters 2000), was apparently alone. Stick Man was an interment, but it is not possible to know if he was interred with a group. Mortuary practices—secondary interment and usual lack of deposition in a cemetery group—are consistent with high mobility over large geographic ranges (Wallthal 1999).

Characterization

In summary, WST people were mobile foragers who created a diverse array of composite implements to use during a highly patterned but widely ranging subsistence round that emphasized big game. This patterned behavior, their longer residence time in base camps, and use of both residence camps and fi eld camps have earned them the label “collector- like foragers” (Ames 1988; Chatters 1985).

Old Cordilleran Tradition

In this discussion, we treat Carlson’s (1996a, 1998) Pebble Tool and Micro-blade traditions as manifestations of the same or closely related technologies within Butler’s (1961) “Old Cordilleran” Tradition. Both include foliate bi-faces, cobble tools, an assortment of fl ake tools, and a maritime or riparian focus. Microblades appear as an alternative technology added to the reper-toire after around 9900 cal B.P. (see Fedje et al. 2008), although the lack of microblades in the earliest components could also be due to sampling issues

TAM Bousman 13106 complete.indd 43TAM Bousman 13106 complete.indd 43 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 8: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

44 Chatters, Hackenberger, Prentiss, and Thomas

(Prentiss and Clarke 2008). Bone implements occur, but typically only as ex-pedient tools. So defi ned, OCT assemblages have been found throughout the Northwest, extending southward from Glacier Bay, Alaska, and the Fraser–Chilcotin confl uence in British Columbia. As we will show later, they fi rst appeared in the north before 10,600 cal B.P., moving progressively southward and inland over a period of 1600 years (table 3.2).

In addition to the tradition name, by which they are known in much of British Columbia, OCT assemblages have been called the Moresby Tradition on Haida Gwaii (Fedje et al. 2008); North Coast Microblade in southeast Alaska (Matson and Coupland 1995); Olcott in western Washington (Butler 1961); and Cascade (Leonhardy and Rice 1970), Vantage (Nelson 1969), and Okanogan phases (Grabert 1968) in eastern Washington, Oregon, and parts of Idaho. Because of broad regional variations in assemblage characteristics, poor site and faunal preservation in heavily forested settings, and interest in contrasting OCT with the WST in a sequence of adaptations, we focus our at-tention on the characteristics of what we hereafter refer to as the “southern interior” manifestations of this tradition, from the area known as the Colum-bia Plateau and northwestern Great Basin.

Technologies

OCT lithic and bone technologies are simpler and food processing technol-ogies more complex than in the WST. Lithic assemblages are dominated by foliate projectile points and bifacial knives, drills, small end and side scrapers,

Table 3.2. Earliest radiocarbon ages of Old Cordilleran and NW microblade components in the Northwest.

Site 14C Age Range Calibrated Age

cal B.P.a Reference Map No.

Groundhog Bay 2 9292 ± 68 10,683–10,303 Ackerman 1968 20Hidden Falls 9060 ± 260 10,553–9785 Davis 1989 22On- Your- Knees Cave 9150 ± 50b 10,380–10,238 Dixon 2008 21Moresby Tradition 8750 ± 60 9888–9627 Fedje et al. 2005 5FlRQ–013 8770 ± 60 9905–9664 Burford et al. 2008 27Landels 8400 ± 90 9516–9304 Rousseau 1991 31Namu 9000 ± 140c 10,367–9888 Carlson 1996b 15Bear Cove 8020 ± 110 9024–8659 Carlson 1979 40Milliken 9000 ± 150 10,369–9800 Borden 1960 23Glenrose Cannery 8150 ± 250 9410–8774 Matson 1996 33Indian Sands (OR) 8250 ± 80 9400–9094 Moss and Erlandson 1996 32Cascadia Cave 7910 ± 280 9089–8420 Newman 1966 42Five- Mile Rapids 8090 ± 90 9189–8778 Butler and O’Connor 2004 39Bob’s Point 7600 ± 110 8540–8326 Minor and Toepel 1986 46Box Canyon 6820 ± 70 7703–7586 Burtchard 1981 53Ash Cave 7940 ± 150 8993–8608 Butler 1962 41Kennewick Mand 8410 ± 40 9489–9417 Chatters 2000

Taylor et al. 199830

Lagoon Site 7320 ± 90e 8277–8016 Draper 1986a 47Plew Site 7730 ± 120e 8633–8394 Draper 1986b 45Hangman Creek 7290 ± 60 8156–8027 Stan Gough pers. comm. 2009 49Marmes Rockshelter 7840 ± 150e 8971–8458 Rice 1969

Sheppard et al. 19873

45WT 2 7300 ± 180 8314–7966 Nance 1966 48Thorn Thicket 7710 ± 180 8766–8336 Sprague and Combes 1966 44Stockoff 7660 ± 780 9441–7742 McPherson et al. 1981 4310NP453 7980 ± 40 8980–8778 Ridenour 2006 25Kirkwood Bar 7100 ± 60f 7980–7856 Reid and Chatters 1997 52Bernard Cr. Rockshelter 7250 ± 80 8162–8005 Randolph and Dahlstrom 1977 5035JE51B 7035 ± 56 8016–7979 Schalk 1995 51

a. One- sigma range, determined using Calib5.b. Dixon (2008) also provides an estimated date of 9200 B.P. for human remains from the site, which is not signifi cantly different from the date given here.c. Carlson (2008) states that the earliest leaf- shaped projectile point from Namu dates to approximately 9000 radiocarbon B.P., placing this compo-nent fi rmly in the OCT. Earlier dates of up to 9700 rcy B.P. on deposits lacking projectile points or microblades could push the age of the OCT on the Northwest Coast considerably further back in time.d. Reference here is to the projectile point embedded in the man’s pelvis. It is arguably not culturally associated with him.e. Shell date.f. Site dates to the time of the Mt. Mazama eruption, making a younger date of 6740 ± 50 B.P. for the earliest occupation stratum the most likely initial age for the site.

TAM Bousman 13106 complete.indd 44TAM Bousman 13106 complete.indd 44 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 9: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

45 Pacifi c Northwest

gravers, and cobble tools (fi g. 3.3). Microblades and microcores are some-times found (e.g., Daugherty et al. 1987a, 1987b; Jaehnig 1984; Munsell 1968). Bifaces often far outnumber other tool categories. Despite a bifacial technology, projectile points were often produced by pressure- trimming fl ake blanks rather than through bifacial percussion reduction. Scrapers and grav-ers tend to be smaller and more formalized than those of the WST. Tools made from cobbles, including choppers, plane- like unifaces, and large spall knives, account for a much higher proportion of the tool inventories than in the WST. Cobbles, systematically reduced in a manner that often produced polyhedral- like cores, account for a high proportion of the lithic material used by this tradition throughout most of its range. Food- grinding tools, which include manos, milling stones, and edge- ground cobbles, are much more com-mon, being found in most assemblages from the southern interior. In the collection from the Granite Point Site, for example, grinding tools comprise nearly 5% of the total tool assemblage in the early OCT component, but only .02% in the WST component. Cobble tools (choppers, spalls, scrapers) make up 16% and 1.6%, respectively (see Leonhardy 1970).

Bone implements are less common, less diverse, and take different forms than in the WST assemblages. Usually, only a few simple, splinter awls and antler wedges are found, but Granite Point and the Marmes and Bernard Creek rockshelters produced a broader array (Hicks 2004; Leonhardy 1970; Randolph and Dahlstrom 1977). Needles remain in the inventory, but are much larger and occur along with simple, splinter awls, as if the awls were used to create openings for needles to thread. Small barbs from composite fi shhooks were found at Granite Point and Bernard Creek; items the excava-tors identifi ed as “leisters” came from the latter site. In general, the inventory indicates far fewer composite tools. Those that do occur—the fi shhooks and leisters—are associated with fi shing.

In addition to milling stones, other food processing technologies were much better developed in the OCT. Earth ovens fi rst appear around 8500 cal B.P. at the Hannavan Creek Site in the southern Willamette Valley (Cheatham 1988), after the OCT was well established in that region (Newman 1966). A “roasting pit” is also reported from site 35JE49 in central Oregon at around 9100 cal B.P. This technology remained a part of the OCT subsistence strategy throughout

Figure 3.3. Stone and bone tools common in the Old Cordilleran Tradi-tion. Included are a microblade core (a), end scraper (b), fi shhook barb (c), metapodial awl (d), edge- ground cobble (e), large biface or knife (f), foliate projectile point (g), drill (h), and cobble chopper (i). Items a, h, and i are from 45D0373; d is from 450K420 (Chatters 2003); f is from Tryon Creek (Hackenberger and Thompson 1995); g is from the middle Columbia River; and c is from Bernard Creek Rockshelter (see Randolph and Dahlstrom 1977). Draw-ings by Claire Chatters (c) and Sarah Moore (all others).

TAM Bousman 13106 complete.indd 45TAM Bousman 13106 complete.indd 45 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 10: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

46 Chatters, Hackenberger, Prentiss, and Thomas

its existence (Thoms 1989). A more common food processing method appears to have been stone boiling. Small, thermally altered rocks are common at OCT sites, often appearing in clusters, as they did at the Plew Site, seemingly having been dumped from baskets after being used for cooking (e.g., Draper 1986b). Although foods were more heavily processed during the OCT, there is no evidence, beyond a single grass- lined pit in at Marmes Rockshelter, of food storage as an important component of the resource management strategy (Chatters 1995).

As food processing technologies improved, people stopped using their teeth as tools. The earliest well- preserved human skeletal remains attributable to the OCT, found at the Braden (Harten 1980), Demoss (Pavesic 1985), and Clark Fork River (Pennefeather- Obrien and Strezewski 2002) sites and dating around 6500–7000 cal B.P., lack the extreme anterior tooth wear and wear on the lingual aspect of lower incisors seen in the WST skeletons (notes of the senior author).

Settlement Strategy

OCT sites are confi ned to mountain ranges and the major river corridors (Chatters and Pokotylo 1998). There is almost no evidence of human activity in the open Columbia Basin during the OCT (Chatters 1982; Greene 1975). Whereas the WST showed a strong tendency for site reuse, the OCT is charac-terized by small sites seemingly used once for perhaps only a few weeks, then abandoned, never to be revisited. Some sites were repeatedly occupied, but these are the exception. OCT sites can be found scattered along both banks of the Columbia and Snake Rivers, wherever surfaces of a suitable age oc-cur. Consistently, they include small concentrations of thermally altered rock, lenses of mussel shells, small collections of cobble tools and fl akes, and a few formed tools (Chatters 1995). Evidence has been found for small dwellings, including small, shallow, semi- subterranean houses at 35JE51B (Schalk 1995) and the Lagoon Site (Draper 1986a), and a surface hut marked by a partial circle of stones at Plew (Draper 1986b). A similar structure- like feature was found in an OCT context at the McCallum site in southern British Colum-bia (Lepofsky and Lenert 2005). The quasi- logistical strategy seen during the WST disappears. In separate analyses of Columbia Plateau assemblages, Bense (1972) and Ames (1988) found no tendency for sites to cluster functionally. There is also no apparent tendency for seasonal habitation patterns, except perhaps for moves between upland and lowland habitats. OCT people during the time before 7500 cal B.P. were highly residentially mobile and tended to engage in many of the same activities at each residence camp, regardless of location or season.

Subsistence

Plant and animal exploitation underwent major changes between the end of the WST and establishment of the OCT in the southern interior. Increased frequencies of grinding stones and the appearance of roasting ovens are evidence for greater dependence on plant foods. By at least 7000 cal B.P., groups in the Rocky Mountains at the DeMoss and Clark Fork Sites were consuming so much processed carbohydrate that it led to extremely high rates of dental caries and antemortem tooth loss (notes of the senior author; Pennefeather- Obrien and Strezewski 2002).

Faunal exploitation patterns are nearly the reverse of those seen for the WST. Review of faunal assemblages from 13 OCT assemblages dating between 9000 and 7500 cal B.P. shows an emphasis on small mammals, fi sh, and, often, freshwater mussels. Among vertebrate prey, fi sh ranked fi rst in NISP at Five- Mile Rapids (Butler and O’Connor 2004), Box Canyon (Burtchard 1981), Bob’s Point (Minor and Toepel 1986), Bernard Creek (Randolph and Dahlstrom 1977), and Kirkwood (Reid and Chatters 1997); second at Ash Cave (Butler 1961) and Plew (Chatters 1986); third at Hat Creek (Shiner 1961); and were important at Judd Peak Rockshelter (Daugherty et al. 1987a). Rabbits and hares were most frequent at 35JE51B (Schalk 1995), Hat Creek, Lagoon (Chatters 1986), and Plew. Deer were the most important animals

TAM Bousman 13106 complete.indd 46TAM Bousman 13106 complete.indd 46 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 11: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

47 Pacifi c Northwestonly at Marmes (Gustafson 1972), Judd Peak, and Ash Cave—all of them, interestingly, rockshelters. Mussels often far outnumber vertebrate remains at these sites; they are a near- ubiquitous component of OCT archaeofaunas

The aquatic focus of OCT subsistence is evident from the two sites out-side the southern interior that have produced good faunal collections. Indian Sands is a shell midden on the Oregon coast (Moss and Erlandson 1996). Stickleback and salmon dominated the faunal assemblage at the Milliken Site on the Fraser River (Matson 1996).

Mortuary Practices

Burials dating to the earlier OCT have been found at six sites, all of which contain multiple individuals, which may be seen as an indication of a pattern of cemetery establishment characteristic of Archaic cultures. Marmes Rock-shelter is the only one of these that predates 7500 cal B.P., however, and the earliest two skeletons at that site (Burials 1 and 2) are the scattered remains of individuals who were not interred. Marmes burials that postdate 7500 cal B.P. are primary interments of adults and children (Breschini 1979; Rice 1969). Clark Fork is an inadvertently discovered pair of skeletons for which con-text information has been lost (Pennefeather- Obrien and Strezewski 2002). The remaining sites are part of the Western Idaho Archaic Burial complex, in which defl eshed and sometimes cremated remains of multiple individuals were interred with caches of outsized and exotic goods during communal burial rituals (Pavesic 1985). While superfi cially similar to the earliest skeletal remains at Marmes, which date to the early WST, this complex postdates the Marmes cremations by some 5000 years.

Characterization

OCT occupants of the southern interior were highly residentially mobile for-agers who lived along major rivers, subsisting primarily on fi sh, freshwater mussels, small mammals, and plant foods. Although they used both lowlands and montane environments, their record shows no sign of seasonal or geo-graphic variability in behavior. They possessed a more advanced food pro-cessing technology, which may have eliminated the need for using their teeth as tools. Despite the heightened mobility, the concentration on plants and small prey, along with evidence in the use of cemeteries of an increased sense of territoriality, is consistent with their inclusion in the Archaic.

The Process of Change from Paleo indian to Archaic

The Archaic OCT replaced the WST in the southern interior of the Pacifi c Northwest between 9500 and 9000 cal B.P. Regional specialists have long seen the two as ancestor–descendant components of the Pioneer Period (e.g., Ames et al. 1998; Daugherty 1962; Davis 2001; Leonhardy and Rice 1970: Matson and Coupland 1995), but the foregoing review indicates otherwise.

The two traditions share many functional types of stone tools and a few bone implements, and sometimes occur in the same sites. Site sizes are small for both, and settlement strategies were mobile. These characteristics, how-ever, can be found in nearly all early traditions in the Americas and across much of Eurasia over a span of tens of thousands of years. The differences, however, far outnumber the similarities, beginning at the most basic level of technology (table 3.3). OCT, throughout its range, shows an emphasis on cobbles as sources of tool material, but also sometimes includes microblades; WST does not. WST projectile tips were bifacially reduced by percussion and did double duty as knives; OCT projectile tips were often pressure fl aked on fl ake blanks. Separate bifaces were prepared as knives. The stone and bone technologies of the WST are more complex than those of the OCT, includ-ing many tools—crescents, bolas, bone foreshafts, bone atlatl spurs, and fi ne needles—that did not persist into the OCT. Most of these are parts of com-posite implements, indicating a high degree of planning depth that the OCT lacked. WST people were somewhat logistically mobile; mobility during the OCT was strictly residential. WST folk made extensive use of all parts of the landscape in a highly patterned manner; OCT bands clung to rivers and high

TAM Bousman 13106 complete.indd 47TAM Bousman 13106 complete.indd 47 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 12: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

48 Chatters, Hackenberger, Prentiss, and Thomas

mountains, doing much the same thing everywhere they went. WST subsis-tence, although rich in species, emphasized large game and made little use of plants; fi sh, shellfi sh, small mammals, and plant foods were the staples of the OCT. The composite tools the OCT did employ were fi shing oriented; ab-sent from the WST, these mark another discontinuity between the traditions. Food processing technology, in addition to making complex starches acces-sible, freed OCT participants from using their mouths as implements.

The differences are indeed deep, reaching all aspects and levels of behavior, but what stand out most starkly as counter to a culturally genetic relationship (sensu O’Brien and Lyman 2000) are the approaches to lithic reduction and the difference in the complexity of both tool forms and land use. In an ances-tor–descendant relationship, we would expect to see continuity at the basic level of technology because it is likely to be learned earliest in life. Here we see no such continuity. A transformation from a complex functional adaptation, with many specialized forms of tools and activities, to a simple one lacking such specializations does not make evolutionary sense. An evolutionary tra-jectory toward increasing specialization is more likely to continue in the same direction than to abruptly revert to generalized behavior. Faced with environ-mental stress and a generalized competitor, a specialized form is likely to be selected against. What this sequence seems to indicate is not a transition from one way of life to another, but a replacement of one lifeway by another—an ethnic replacement (using the anthropological, not lay, defi nition of “ethnic”).

Three other lines of evidence seemingly support this interpretation. These are basketry styles in the northwestern Great Basin, where the WST/OCT sequence can be seen; the physical characteristics of the human beings them-selves; and the time–space distribution of components of the two traditions.

Basketry Styles

Connolly and Barker (2004) recently conducted a survey of radiocarbon- dated basketry styles in the northern Great Basin, which is the southern extremity of our study region. Looking at designs in twined sandals and basket fragments, they found a distinct break in manufacturing techniques at ca. 9000 cal B.P. Before that date, only simple twined basketry was produced, but a wide array of decorative conventions characterized the basketry thereafter. The change in sandals is even more striking. Except for continuity at the most basic structural levels, all other aspects of sandal design changed abruptly at ca. 9300 cal B.P. Three sandal styles—Fort Rock, Open Twined, and Spiral Weft—are recog-

Table 3.3. Comparison of Western Stemmed and Old Cordilleran Traditions in the southern interior of the Pacifi c Northwest.

Dimension Western Stemmed Old Cordilleran

Land use Widespread use of entire region, including upland plateaus

Restricted to rivers, high mountains

Environmentally patterned tool use No environmental patterning of tool useSettlement Seasonal round with different activities per season No evident seasonal or geographic variation

in activityHigh frequency of site reuse Low frequency of site reuse

Mobility High frequency but longer duration than OCT High frequencySemi- logistical Strictly residential

Technology High degree of planning depth (many composite tools)

Low degree of planning depth (few composite tools)

Lithics not cobble based Cobbles primary tool sourceDiverse stone and bone toolkits Simple stone and bone toolkitsFood processing simple, limited to rare grinding implements

Food processing complex, included grinding, boiling, pit roasting

Use of teeth as tools Teeth not used as toolsSubsistence Focus on big game; little use of fi sh Focus on fi sh, mussels, small game

Little use of plants Increased emphasis on plantsMortuary Secondary cremation, isolated abandonment, and

intermentAbandonment initially, later both primary and secondary cemeteries

Basketry Simple twined throughout Wide array of decorative twined formsFt. Rock sandals Open Twined and Spiral Weft sandals

Human morphology Strong prognathism/facial forwardness Faces orthognathic

TAM Bousman 13106 complete.indd 48TAM Bousman 13106 complete.indd 48 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 13: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

49 Pacifi c Northwest

nized by the region’s specialists (fi g. 3.4). Fort Rock style sandals were made with a twined toe fl ap and open heel; they were held on by cords wrapped around the ankle. Open Twined and Spiral Weft type sandals have heel pock-ets and were tied on with a cord that ran from the toe to the ankle and was drawn tight through a series of loops that reached from the edge of the sole to the top of the foot. Spiral Weft sandals have an open toe, whereas Open Twined sandals have a toe fl ap of loose wefts. Fort Rock sandals all date before 9200 cal B.P. Spiral Weft and Open Twined forms only postdate 9400 cal B.P. Noting that, because of its plasticity, basketry style has long been recognized as an indicator of ethnicity in the western United States, Connolly and Barker suggest that 9000 cal B.P., which is a time of widespread cultural changes, may mark a time of “population change” in the northern Great Basin (2004:250).

Human Characteristics

A geographic displacement of one ethnic group by another need not be marked by widespread genetic changes, culture being a learned phenomenon rather than an inherited one. That caveat notwithstanding, the few human skeletons that have been dated to the Early to Early–Middle Holocene in the Pacifi c Northwest do exhibit notable differences in cranial morphology. In the southern interior of the Pacifi c Northwest, only two individuals predating 9000 cal B.P. are intact enough to provide some information about morphol-ogy. These are Kennewick Man (9400 cal B.P.; Chatters 2000; Powell and Rose 1999; Tayor et al. 1998) and Stick Man (9100 cal B.P.; Chatters et al. 2000), both from Washington. The oldest remains from the southern interior that postdate 9000 cal B.P. and are intact enough for comparison come from two multiple interments in Idaho: Braden (6600 cal B.P.; Harten 1980) and DeMoss (6800 cal B.P.; Green et al. 1986). Measurements were all made by the senior author using standard osteometric equipment.

Cranial morpholometric data could be obtained from Kennewick, Stick, two males each from Braden and DeMoss, and one female from Braden. To compare the crania, we used four simple measures: cranial index, upper facial index, height index, and facial forwardness index. The fi rst two are standard measures (Bass 1987). Height and facial forwardness measures were devised for this analysis to account for the relative incompleteness of most specimens. The height index is computed as 100 × the bregma radius (Howells 1973) divided by the mean of maximum length and maximum breadth. For the computation of facial forwardness, nasion radius was substituted for bregma radius in the equation. The DeMoss female was excluded from computations of mean cranial index because of obvious tumpline deformation, which gave her an extremely elongated, narrow, saddled neurocranium (cranial index 66.7). Results are presented in table 3.4.

Figure 3.4. Sandal styles from the southern interior of the Pacifi c Northwest (Northern Great Basin), showing the sharp distinction in design at 7200–2400 cal b.c. (from Connolly and Barker 2004: fi g. 1, by permission from Oregon State Museum of Anthropology.

TAM Bousman 13106 complete.indd 49TAM Bousman 13106 complete.indd 49 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 14: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

50 Chatters, Hackenberger, Prentiss, and Thomas

Although the sample is painfully small, WST individuals tend to differ from their OCT counterparts in having longer, narrower, higher neurocrania, and a narrower, more forward- positioned face. The only difference that is signifi -cant, however, is the degree of facial forwardness (fi g. 3.5), for which mean indices of the two groups do not intersect at 3 standard deviations. Facial forwardness is the most robust, continent- wide difference between Paleo-indians and later American peoples (e.g., Chatters 2000; Jantz and Owsley 2001). These differences lend support to the impression that the changes tak-ing place in the southern interior of the Pacifi c Northwest at around 9000 cal B.P. were the result of a population replacement. They must be viewed with extreme caution, however, and should not by themselves be considered prima facie evidence of such an event. The sample is simply too small and the time elapsed between the two sets of remains too great for this to be consid-ered primary evidence.

Traditions in Time and Space

The chronology of OCT entry into the Pacifi c Northwest provides further sup-port for the idea of an ethnic replacement. Figures 3.6a through 3.6c map the distribution of dated WST and OCT sites in the Pacifi c Northwest at >10,000, 10,000–9000, and after 9000 cal B.P. At 10,000 cal B.P., only WST sites oc-curred in the southern interior, while OCT assemblages are documented only

Table 3.4. Craniometric indices for Western Stemmed Tradition and Old Cordilleran Tradition groups. Number in parentheses is the N for the measure.

Measure WST OCT

Cranial index 72.2 ± 1.9 (2) 75.0 ± 2.7 (4)Upper facial indexa 55.6 (1) 52.6 (1)Height index 77.8 ± 4.8 (2) 74.9 ± 3.1 (2)Forwardness index 63.0 ± 0.3 (2) 59.5 ± 1.1 (2)

a. Only Kennewick and Braden “unnumbered” skulls could be used for this measure.

Figure 3.5. Marked difference in facial forwardness between

Late Paleoindian (Kennewick) and early Archaic (Braden) skulls, overlain in the Frankfurt plane.

TAM Bousman 13106 complete.indd 50TAM Bousman 13106 complete.indd 50 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 15: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

51 Pacifi c Northwest

along northern Northwest Coast, south as far as the mouth of the Fraser River. By 9400 cal B.P., OCT could be found in the interior of British Columbia and the Pacifi c Coast as far south as southern Oregon. It also began to appear at points along the Columbia River in the westernmost part of the southern inte-rior, such as the projectile point embedded in Kennewick Man’s pelvis (Fagan 1999; Chatters 2000). By shortly after 9000 cal B.P., it had displaced the WST in all but the southeasternmost part of its range. This time- transgressive dis-tribution is indicative of a slow expansion of a cultural tradition south from eastern Beringia, millennia after the fi rst peopling of the continent (Chatters 2010b).

The Evidence from Beech Creek

Although the record of technological, settlement, land use, and physical anthropological change can be interpreted as the replacement of one cul-ture by another, the limited archaeological record from the time period be-tween 9500 and 8800 cal B.P. (tables 3.1 and 3.2) leaves open the possibility

Figure 3.6. The distribution of OCT and WST occupations of the Pacifi c Northwest at >10,000 B.P., 10,000–9000 B.P., and after 9000 B.P.

TAM Bousman 13106 complete.indd 51TAM Bousman 13106 complete.indd 51 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 16: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

52 Chatters, Hackenberger, Prentiss, and Thomas

that the exchange of ideas and intermarriage between neighboring groups, stretching over the span of a few centuries, could account for the trans-formation. The issue could be resolved with a larger sample of individual occupations from each tradition dated to this critical time interval or with a continuous, discrete stratigraphic sequence containing assemblages of the two traditions, either with or without transitional forms. In preparing this article, we sought sites that contained both traditions in a clear stratigraphic sequence; unfortunately, the quest was fruitless. All sites that contained both traditions in apparently continuous sequence had signifi cant limita-tions. Five- Mile Rapids, seemingly one of the best candidates because of deep stratifi cation and rapid deposition, had been excavated before some of the issues discussed here were being considered and was reported in a man-ner that makes distinguishing individual assemblages extremely diffi cult (Cressman et al. 1960). High water table prevented recent excavations at the site from reaching strata that predate 9100 cal B.P. (Butler and O’Connor 2004), into the WST levels. Marmes Rockshelter, with a purported continu-ous sequence (Hicks 2004), was excavated in broad stratigraphic units. The transition between WST and OCT is included in Unit III, which spans the period from around 9500–7500 cal B.P. Not surprisingly, both WST and OCT elements are found in that stratum. Paulina Lake exhibited some of the same problems. The entire pre- 7500 cal B.P. deposit was only 50 cm thick and lacked stratigraphic breaks of any kind. Connolly (1999) was, therefore, forced to arbitrarily designate three subunits within this series, the middle of which, this time, contained both WST and OCT elements. Rock Island Over-look was another shallow deposit in which assemblages were commingled; it has never been fully analyzed and reported (Valley 1975). Only Granite Point (Leonhardy 1970), which has been discussed briefl y above, contained discrete WST and OCT assemblages predating 7500 cal B.P., but they are widely separated by a fl uvial series of seeming long duration. There seemed no suitable candidate for clarifying the Paleo indian–Archaic interface. Then, in the fall of 2007, two of us (Chatters and Prentiss 2010) obtained a con-tract to analyze lithic artifacts from the Beech Creek Site (45LE415) in west-ern Washington.

The Beech Creek Site

Beech Creek is an open site on the fl oor of the broad upper Cowlitz River Valley near Packwood, Washington, only 16 kilometers west of the crest of the Cascade Range. Located at an elevation of only 317 meters, it occupies a small, fl at- topped ridge or terrace remnant beside deeply incised Beech Creek, overlooking the modern Cowlitz River fl oodplain. Rick McClure and Cheryl Mack of the Gifford Pinchot National Forest conducted excavations at the site in 1992 and 2006, recovering more than 40,000 stone artifacts.

The site is discretely stratifi ed in a series of very poorly sorted layers of what appear to us to be either colluvium from the adjacent mountain slopes or alluvial fan sediments from Beech Creek, laid down before the stream incised its present- day canyon. Bioturbation, largely from tree roots, is ex-tensive. Artifacts were all found in Stratum IV, a massive layer that averages approximately 120 cm thick and is capped by tephra from the 470 cal B.P. eruption of Mount St. Helens (McClure 1992). Sedimentation during the de-position of Stratum IV appears to have been slow but continuous. Deposi-tion of the stratum ceased long before the recent volcanic eruption. The tool assemblage consists of a wide variety of bifacial and unifacial implements made from cryptocrystaline silicates, microcrystalline volcanics, and obsid-ians, plus a small number of cobble implements of mostly volcanic origin. Unfortunately, little bone is preserved and no domestic features were found. Projectile point styles, which include Windust stemmed, Cascade leaf- shaped, and large, corner- notched and barbed specimens resembling tools from British Columbia’s Nesikep Tradition, indicate occupation at the site probably ended before 7500 cal B.P. (Chatters and Prentiss 2010). The vertical distribution of artifacts is more or less even, with some zones of greater or lesser frequency but no patterning of frequency modes. At fi rst look, this assemblage appeared

TAM Bousman 13106 complete.indd 52TAM Bousman 13106 complete.indd 52 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 17: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

53 Pacifi c Northwestto be beset by the same problems as Marmes Rockshelter, Rock Island Over-look, and Paulina Lake. Analysis proved otherwise.

Methods of Analysis

Twenty- fi ve 1- meter square units were excavated during 2006 as part of a data recovery effort associated with the planned sale of federal land. McClure and Mack asked us to conduct technological analysis of debitage from the four units with the thickest Stratum IV and largest quantities of lithics. We also conducted stylistic and functional analyses of all 306 tools from the 2006 excavations, including utilized fl akes, and a technological analysis of all cores. Debitage analysis included aggregate analysis and simplifi ed reduction stage categorization, both by material. Tools were identifi ed to traditional formal classes; each employable unit (per Knudson 1983) on each tool was analyzed macroscopically and microscopically for wear patterns. Projectile points were classifi ed into regional types based on its plan view, blade cross- section, fl ak-ing patterns, and edge treatment.

Once the analysis was complete, we sought evidence for intra- assemblage variability. It is important to note that the above- described analyses were con-ducted blind, without any information but the lithics themselves and their horizontal and vertical provenience. We submitted the data tables from this analysis to McClure and Mack before seeking any information about sedi-mentary context that might be used to distinguish vertical or horizontal pat-terning. After a principal components analysis of the debitage from the four sample units indicated the presence of a zone containing higher frequencies of early- stage core reduction debitage, we asked if there were any horizons in which large stones consistently occurred. In reply, Mack provided data show-ing there had been such a horizon in all but fi ve of the 25 units. We used this manuport- strewn horizon to divide the assemblage into three analytic zones: I, above the stony horizon; II, the stony horizon itself; and III, all material from below the stony horizon. Looking at the distribution of projectile point styles across these zones, we found that all of the Western Stemmed projectile points but one were from Zone III. The odd point out was from 10–20 cm above the Zone II/III boundary, in Zone II. Conversely, all but one of the Cascade- style points from the OCT came from Zones I and II. The exception was a tiny basal fragment from 20–30 cm below the stony horizon. This separation based on the stony horizon had resulted in discrete WST and OCT components in a site with no apparent hiatus in archaeological deposition.

Comparison of Assemblages

In our comparison of the three assemblages, each of which consisted of be-tween 94 and 113 tools, we sought to determine if the assemblages of the traditions exhibited differences that were comparable with those described in table 3.3 for the region as a whole and to determine if they offered evidence that might support or refute the idea of ethnic replacement. Differences in settlement and mobility patterns, activities performed at the site, use of indi-vidual tool classes, and passive style were the foci of our analysis. We found a consistent pattern of difference, often statistically signifi cant, across nearly all variables considered. The only measure that produced a signifi cant difference between Zones I and II was the proportion of early- stage reduction debitage, which was the variable that initiated identifi cation of the zones.

Settlement and Mobility

Consideration of the region- wide record (table 3.3) indicates that WST folk were somewhat logistically mobile and “collector- like” in their mobility strat-egy, using both residence camps and fi eld camps, whereas OCT people were residentially mobile. Core types, intensity of tool utilization and rejuvenation, and levels of expediency, as indicated by the ratio of expedient tools to curated tools in the Beech Creek assemblages, all support this characterization. Cores in Zone III were primarily spent, bipolar cores; those in Zones I and II were nearly all polyhedral- like cobble cores. The difference is statistically signifi cant. Zone III occupants made much more intensive use of their toolkit (table 3.5).

TAM Bousman 13106 complete.indd 53TAM Bousman 13106 complete.indd 53 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 18: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

54 Chatters, Hackenberger, Prentiss, and Thomas

Utilized fl akes in Zone III have 18% more employable units (EU) per item than either Zone I or II. Retouched tools have more EU per tool, a higher percentage of them are worn, and bifaces were more heavily worn in Zone III than in the later zones. In Zone III, all Stage 4 bifaces (with pressure retouch) other than projectile points had edge wear; few in the other two zones were worn. Worn tools with their edges showing signs of multiple resharpening episodes were common in Zone III, rare elsewhere. Small debitage, which represents tool fi nishing and rejuvenation, is signifi cantly more abundant in Zone III than in each of the other two zones. As mentioned earlier in this pa-per, the WST component of this site showed a higher ratio of expedient tools to curated ones: .98 for Zone III and .79 for Zones I and II. All of these mea-sures bespeak a longer duration of residency each time Zone III people made camp at Beech Creek. It was more of a base camp during the WST occupation, a short- term residence camp when visited by OCT people.

Use of Tool Classes

Tool use frequencies show differences in the function of what we interpret to be the same tool forms, as well as emphasis on different activities between OCT and WST occupations. In Zone III, microscopic wear patterns on utilized fl ake edges show an emphasis on scraping actions. These fl akes and retouched scraping tools were used for working both hard (bone, wood) and soft (skin) materials. Stage 4 bifaces (knives) were used for slicing and chopping both soft and tough materials—that is, for such activities as skinning, meat slicing, and heavy butchering, and perhaps cutting hardened skins. Utilized fl akes in Zones I and II were used primarily as cutting tools, perhaps assuming many of the functions held by knives in Zone III. Scraping tools were used primarily on soft materials, mainly for skin working. Knives were used only for slicing tough materials. In general, tools were used for a much wider range of activi-ties during the WST occupation, another indicator of longer, base camp–like residency. OCT folk concentrated on producing tool blanks and processing meat and hides while at the site; WST inhabitants performed a much wider array of activities, including tool fi nishing and repair, and working wood, bone, and fresh and dried skins.

The most striking difference in tool utilization is seen in the projectile points, which show almost mutually exclusive patterns of edge damage and breakage. All but one blade among the Western Stemmed projectile points exhibits edge rounding and “greasy” edge- and- fl ake- scar polish of the type produced by slicing meat. All of the worn points have been resharpened over this edge wear, leaving small islands of polish. None of the Cascade Point blades show any edge wear. All of the fi ve broken WST points have bend-ing fractures; none are impact fractured. Of ten non- fi re- fractured Cascade points, eight are impact fractured and only two show bending fractures. The WST points were used both as projectiles and as knives; OCT points served only one function.

It is tempting to see a relationship between the use of projectiles as meat- cutting knives and WST peoples’ paramasticatory use of their teeth. Perhaps with limited food processing technology, they ate much in the manner ob-served among Eskimo and Aleut people, by grasping meat between the teeth and one hand while cutting off a bite with the other.

Passive Style

Ames (2000) makes a distinction between active and passive styles. Active style refers to differences in the form of functionally similar implements, garments, or vessels that we assume to have had conscious social meaning.

Table 3.5. Intensity of tool use between WST (Zone III) and OCT (Zones I and II) at the Beech Creek site.

Zone EU/Utilized Flakes EU/Tools Worn Tools/Total Tools Worn Bifaces/Total Bifaces

I 1.29 .80 .88 .10II 1.28 .92 .96 .22III 1.44 1.14 1.09 .47

TAM Bousman 13106 complete.indd 54TAM Bousman 13106 complete.indd 54 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 19: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

55 Pacifi c NorthwestFort Rock sandals were visibly distinct from Spiral Weft sandals, and would thus have been useful signals of social group membership. We assume the large, broad Western Stemmed projectile points and narrow, thick, usually serrated- edged willow leaf Cascade points found at Beech Creek also con-veyed such meaning. Passive style refers to differences in manufactured ob-jects that are the result of “rote social learning” (Bettinger et al. 1996). We previously referred to basic tool technology as an indicator of continuity be-tween one broad cultural pattern and its descendant. Such things as fl aking patterns, raw material choices, forms of fl akes selected for blanks, and sizes of selected and produced implements are considered here to be elements of passive style. Patterns of projectile point use are, we believe, another.

Multiple differences that can be considered discontinuities in passive style are found in the Beech Creek collection. WST occupants preferred micro-crystalline volcanic rock over cryptocrystalline silicas for fl ake blanks. They also preferred hard hammer percussion fl akes over biface thinning fl akes and chose larger fl akes with less acute edge angles than selected by the later, OCT occupants. The blanks they selected were larger, as were bifaces they produced at the site. When they produced side scrapers, they made them with less acute angles than their successors did. OCT people preferred smaller, biface thinning fl akes of cryptocrystalline silica and selected more acute edge angles. We have already discussed the extremely different ways they used projectile points. The difference in fl ake blank preference is signifi cant, as are differences in edge wear and breakage on projectile points. These fi ndings provide support for the replacement hypothesis and refute the idea of an ancestor–descendant relationship between the WST and OCT in the southern interior.

Discussion

The change in the interior Pacifi c Northwest between the WST (Paleo indian) and OCT (Archaic) entails all aspects of behavior—land use, settlement pat-terns, mobility strategy, technological content, complexity, and planning depth, and both active and passive style. It entails elements of technology that, within anthropology, are traditionally considered the purview of both men (lithic and bone technology) and women (food processing, basketry). It also is loosely correlated with a change in human craniofacial morphology. We have suggested here that this pattern is consistent with ethnic replace-ment. There are, however, alternative explanations that could account for many, if not most of the cultural changes we have identifi ed.

Many of the changes might be explained as a response to the increasing aridity and warming winters that occurred between 9500 and 9000 cal B.P. The reorganization of OCT settlement and land use patterns from upland ter-restrial environments to the major river corridors, change to fi sh and shellfi sh as mainstays of subsistence, and preference for cobble sources of stone might all be accounted for as a response to the change in the balance of resources from the drying uplands to the more stable rivers (Davis’s [2001] oasis ef-fect; see also Lyman 1992). The change in bone technology from hunting implements toward fi shhooks and leisters would be a manifestation of that new focus. Increasing emphasis on plants might also be interpretable as a response to the aridity- induced loss of large mammalian prey as a source of carbohydrates, although anadromous fi sh could have been a suffi cient re-placement. The abandonment of seasonal differences in subsistence- oriented behavior and the shift from partially logistical to residential mobility might be accounted for as a response to rising effective temperature (sensu Binford 1980) toward the end of the Early Holocene.

Some tool categories—one could say whole technological complexes—were lost at the end of the WST, whereas the OCT ushered in new tools and complexes. We have already mentioned how the introduction of fi shing implements could be a response to resource availability, but what of bolas and crescents? Perhaps they were associated with species or habitats that were lost or declined regionally as the environment dried out, but without knowing the tools’ original uses we are hard pressed to identify what those changes might have been. Perhaps the shift away from using teeth as implements has

TAM Bousman 13106 complete.indd 55TAM Bousman 13106 complete.indd 55 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 20: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

56 Chatters, Hackenberger, Prentiss, and Thomas

a similar explanation: reduction in large ungulate prey and, hence, large pack-ages of tough meat. It could also be a collateral effect of new food processing technologies (Chatters 2008).

Style changes cannot be linked to climatic events, at least in a direct causal sense. Nothing in the shift from upland to riparian land use could account for the replacement of stemmed points and distally broad, leaf- shaped weapon tips by proximally broad, leaf- shaped points. Neither can it explain the re-placement of Fort Rock sandals by distinctly different styles. Likewise, it can-not account for the absence of change in both artifact categories for thousands or years before and after the appearance of the OCT.

Physical changes in the human form are also not explained with simple ref-erence to environmental causes. Sampling error due to small sample sizes and genetic drift in small populations (see Powell 2004) could account for the ob-served differences. Time is also a factor, with compared specimens separated by some 2500–3000 years. Natural selection could have favored less forward faces, or not favored more forward ones as new processing technologies and less- tough meat required lower bite forces in mastication.

The time–space patterning of the OCT in the Northwest is more diffi cult to explain. The only ways to negate this as a support for ethnic movement are to say that we have either incorrectly lumped multiple distinct technologi-cal traditions or that the record is simply too incomplete along the southern Northwest Coast for us to say what technology occupied that area before the OCT is recognized. This alternative is contradicted by the Youngs River Com-plex variant of the WST near the mouth of the Columbia River and in parts of lowland western Washington.

We acknowledge that there is at least one alternative explanation for many of the changes we have noted between the WST and OCT. Such an account-ing, however, requires multiple, often unrelated explanations. Ethnic replace-ment is a single process that can account for all observed differences. Fur-thermore, multiple explanations do not account for changes in passive style, which should be immune to environmental infl uence. Replacement is, there-fore, the more elegant and empirically suffi cient, hence better explanation for the Paleo indian–Archaic shift in the interior Northwest. The presence of change in both male and female products, particularly changes in style, makes it clear that the entire population was involved in the move; intermarriage and acculturation, although they almost certainly occurred, played little part.

The conclusion that increasing aridity did not bring about the subsistence, technological, settlement, and land use differences between the two traditions does not mean that it was not involved in the change from WST to OCT. In fact, we are certain that environmental change played a pivotal role in this event. It is widely acknowledged (e.g. Carlson 1996a; Fedje et al. 2008) that the OCT, as we have defi ned it here, was adapted to maritime conditions; the WST, although not averse to consuming fi sh and shellfi sh, emphasized ter-restrial resources, particularly large ungulates in our region. At 9500 cal B.P., we can envision two neighboring groups in the southern Pacifi c Northwest: the OCT occupying the coast and the WST entrenched in the interior. By this time, the OCT apparently had developed or adopted stone boiling and earth oven processing methods and, therefore, the ability to use complex plant starches. Over the next 500 years, the terrestrial environment along the coast remained stable, while that of the interior, east of the Cascade Range, became more arid and less productive. At the same time, coastal upwelling was reestablished, increasing nearshore marine productivity, which would have included anadromous fi shes that used both coastal and interior streams. Thus conditions improved for marine- adapted coastal peoples, while they de-teriorated for their terrestrially adapted interior neighbors. The adaptation to aquatic resources gave the OCT a selective advantage over the now- weakened WST, who were soon displaced.

The southward immigration of the OCT from Alaska into temperate North America between 10,600 and 9000 cal B.P. was secondary to one or more late Pleistocene migrations that constituted the initial peopling of the Americas (Goebel et al. 2008). Such a later peopling event is seemingly contradicted

TAM Bousman 13106 complete.indd 56TAM Bousman 13106 complete.indd 56 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 21: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

57 Pacifi c Northwestby mitochondrial, Y- chromosome, and nuclear DNA evidence that all Native American peoples, past (Gilbert et al. 2008) and present, derive from a single ancestral population (Tamm et al. 2007). Genetic evidence, however, demon-strates not a single founding population for the Americas, but rather a single source population, which was isolated in Beringia for a thousand to more than 10,000 years before budding off one or more groups that became the Ameri-can founders. Archaeological evidence demonstrates the existence of three distinct stone tool traditions, and thus probably at least three separate ethnic groups and at least as many gene pools, in eastern Beringia prior to the move-ment of the OCT onto the Northwest Coast. Earlier immigrants, from whom the WST folk descended, may have represented one of those traditions, the OCT apparently represented another. In their summary of evidence for the early peopling of the Americas, Geobel et al. (2008) acknowledge that Clovis might represent a later immigration from the same Beringian source popula-tion that gave rise to pre- Clovis peoples. Following this same reasoning, the OCT was apparently a third such event. It is likely that there have been others, the last being the Athabascan expansion that gave rise to the Apache, Navajo, Tlatskanie, Nikola, Hupa, and other groups less than 2000 years ago.

Conclusion

The Paleo indian–Archaic transition in the Pacifi c Northwest is an example of a macroevolutionary event in which a maritime- adapted coastal population with advanced fi shing and food processing technologies took advantage of weakened terrestrial productivity and improved oceanic conditions, and sup-planted their Paleo indian neighbors. The event, recognized archaeologically by the replacement of the Western Stemmed Tradition by the Old Cordilleran Tradition, took place between 9500 and 9000 cal B.P. in the Columbia Plateau and neighboring parts of the Great Basin. It is represented by synchronous changes in land use, settlement patterns, mobility strategies, subsistence em-phases, and lithic, bone, basketry, and food processing technology. The fact that this event is accompanied by changes in human morphology, and both active and passive styles of utensils made by women as well as men, makes ethnic replacement the most likely explanation. No other single process can account for such a complex and pervasive transformation. This event repre-sents a secondary expansion of founding peoples out of the Beringian heart-land.

Notes

1. This measure is useful only for intrasite comparisons due to differences in analysts’ interpretation of tool forms. Ratios for each tradition are, there-fore, not expected to be equivalent. The ratio of expedient to curated tools at Granite Point, based on tallies in Leonhardy (1970), is 3.76 for WST and 1.36 for the pre- Mazama OCT assemblage. At Beach Creek the ratios were .98 and .79, respectively (Chatters and Prentiss 2010).

2. Wilson (2008) found a higher frequency of pygmy rabbit bone in Lind Coulee than reported by Irwin and Moody, leading him to infer that this indicates greater emphasis on small prey. Lind Coulee was, however, pygmy rabbit habitat and the presence of this tiny animal at the site cannot be inferred as cultural. It is more elegantly explained as the equivalent of the mouse and ground squirrel remains that also occur in the site.

3. Butler (2004) reports large quantities of nonsalmonid fi sh bone from the Marmes and Harrison horizons at Marmes Rockshelter. This assemblage was collected from fl oodplain deposits and using a fi ner mesh screen than employed at any other site in the region. The bone is more likely to represent fi sh stranded by fl oodwaters than food remains. Unfortunately, no non- site control was analyzed to enable this inference to be ruled out.

References Cited

Ackerman, R. E. (1968). “The Archaeology of the Glacier Bay Region, Southeastern Alaska.” Report of Investigations No. 44. Pullman: Washington State Univer-sity Laboratory of Anthropology.

———. (1996). “Early Maritime Culture Complexes of the Northern Northwest

TAM Bousman 13106 complete.indd 57TAM Bousman 13106 complete.indd 57 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 22: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

58 Chatters, Hackenberger, Prentiss, and Thomas

Coast.” In Early human Occupation in British Columbia, edited by R. L, Carlson and L. Dalla Bona, 123–32. Vancouver: University of British Columbia Press.

Aikens C. M., D. L. Cole, and R. Stuckenrath. (1977). “Excavations at Dirty Shame Rockshelter Southeastern Oregon.” Tebiwa 4:1–29.

Ahler, S. A. (1971). “Projectile Point Form and Function at Rodgers Shelter, Mis-souri.” Research Series No. 8. Columbia: Missouri Archaeological Society.

Ames, K. M. (1988). “Early Holocene Forager Mobility Strategies on the Southern Columbia Plateau.” In Early Human Occupation in Far Western North America: The Clovis- Archaic Interface, edited by J. Willig, C. M. Aikens, and J. L. Fagan, 325–60. Anthropological Papers No. 21. Carson City: Nevada State Museum.

———. (2000). “Review of the Archaeological Data. In Kennewick Man, complied by F. P. McManamon. National Park Service. http://www.nps.gov/archeology/kennewick/ames.htm.

Ames, K. M., D. E. Dumond, J. R. Galm, and R. Minor. (1998). “Prehistory of the Southern Plateau.” In Handbook of North American Indians, Vol. 12: Plateau, edited by D. E. Walker, 103–19. Washington, DC: Smithsonian Institution.

Ames, K. M., J. P. Green, and M. Pfoertner. (1981). “Hatwai (10NP143): Interim Re-port.” Archaeological Reports No. 9. Boise, ID: Boise State University.

Atwell, R. G. (1989). “Subsistence Variability on the Columbia Plateau.” Unpublished MA thesis, Portland State University, Portland, O.

Barnosky, C. W., P. M. Anderson, and P. J. Bartlein. (1987). “The Northwestern US during Deglaciation: Vegetational History and Paleo climatic Implications.” In North American and Adjacent Oceans during the Last Glaciation, edited by W. F. Ruddiman and H. E. Wright Jr., 289–321. Boulder, CO: Geological Soci-ety of America, Boulder, Colorado.

Bass, W. M. (1987). Human Osteology: A Laboratory and Field Manual, 3d ed. Columbia: Missouri Archaeological Society.

Bedwell, S. F. (1973). Fort Rock Basin: Prehistory and Environment. Eugene: University of Oregon Books.

Bense, J. A. (1972). “The Cascade Phase: A Study in the Effect of the Altithermal on a Cultural System.” Unpublished PhD dissertation, Washington State Univer-sity, Pullman.

Bettinger , R. L., R. Boyd, and P. J. Richerson. (1996). “Style, Function, and Cultural Evolutionary Process.” In Darwinian Archaeologies, edited by H. G. D. Mas-chner, 113–64. New York: Plenum Press.

Binford, L. R. (1980). “Willow Smoke and Dogs’ Tails: Hunter- Gatherer Settlement Systems and Archaeological Site Formation.” American Antiquity 45:4–20.

Borden, C. E. (1960). DjRi3, an Early Site in the Fraser Canyon, British Columbia. Contri-butions to Anthropology, Bulletin No. 162. Ottawa, ON: National Museum of Canada.

Brauner, D. R. (1985). “Early Human Occupation in the Uplands of the Southern Plateau: Archaeological Excavations at the Pilcher Creek Site, Union County, Oregon.” Report on File. Corvallis: Department of Anthropology, Oregon State University.

Breschini, G. S. (1979). “The Marmes Burial Casts.” Northwest Anthropological Research Notes 13:111–75.

Burford, A. K. C., F. Craig, R. N. V. F. Arvecque, and N. Jackman. (2008). “An Early Cordilleran Assemblage from the Nechako- Fraser Basin.” In Projectile Point Sequences in Northwestern North America, edited by R. L. Carlson and M. P. R. Magne, 293–302. Burnaby, BC: Archeology Press, Simon Fraser University.

Burtchard, G. C. (1981). Test Excavations at Box Canyon and Three Other Sites in the Mc-Nary Reservoir. Project Report No. 10. Pullman: Washington State University.

Butler, B. R. (1961). “The Old Cordilleran Culture in the Pacifi c Northwest.” Occa-sional Paper No. 5. Pocatello: Idaho State College Museum.

———. (1962). “Contributions to the Prehistory of the Columbia Plateau: A Report on Excavations in the Palouse and Craig Mountain Sections.” Occasional Paper No. 9. Pocatello: Idaho State College Museum.

———. (1965). “A Report on Investigations at an Early Man Site near Lake Channel, Southern Idaho.” Tebiwa 9(2):1–20.

Butler, V. L. (2004). “Fish Remains.” In Marmes Rockshelter: A Final Report on 11,000 Years of Cultural Use, edited by B. A. Hicks, 319–337. Pullman: Washington State University Press.

Butler, V. B., and J. C. Chatters. (2003). “The Holocene History of Salmon in the Co-lumbia River Basin.” Geological Society of America Abstracts with Program 35:607.

TAM Bousman 13106 complete.indd 58TAM Bousman 13106 complete.indd 58 4/30/12 8:26 AM4/30/12 8:26 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 23: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

59 Pacifi c NorthwestButler, V. L., and J. E. O’Connor. (2004). “9000 Years of Salmon Fishing on the Co-lumbia River, North America.” Quaternary Research 62:1–8.

Carlson, C. C. (1979). “The Early Component at the Bear Cove Site, b.c.” Canadian Journal of Archaeology 3:177–94.

Carlson, R. L. (1996a). “Prologue: Introduction to Early Human Occupation in British Columbia.” In Early Human Occupation in British Columbia, edited by R. L. Carl-son and L. R. Dalla Bona, 3–10. Vancouver: University of British Columbia Press.

———. (1996b). “Early Namu.” In Early Human Occupation in British Columbia, edited by R. L, Carlson and L. Dalla Bona, 83–102. Vancouver: University of British Columbia Press.

———. (1998). “Cultural antecedents.” In Handbook of North American Indians, Vol. 7: Northwest Coast, edited by W. Suttles, 60–69. Washington, DC: Smithsonian Institution.

Carlson, R. L., and M. P. R. Magne. (2008). “Projectile Ppoints Past and Present.” In Projectile Point Sequences in Northwestern North America, edited by R. L. Carlson and M. P. R. Magne, 1–10. Burnaby, BC: Archeology Press, Simon Fraser University.

Chatters, J. C. (1982). “Prehistoric Settlement and Land Use in the Dry Columbia Basin.” Northwest Anthropological Research Notes 16:125–47.

———. (1985). “Evolution in Hunter- Gatherer Adaptive Strategies on the Columbia Plateau.” Manuscript in possession of the author, Bothell, WA.

———. (1986). The Wells Reservoir Archaeological Project, Washington, Vol. 1. Central Washington Archaeological Survey No. 86–6. Ellensburg: Central Washington University.

———. (1995). “Population Growth, Climatic Cooling, and the Development of Col-lector Strategies on the Southern Plateau, Western North America.” Journal of World Prehistory 9:341–400.

———. (1996). Paleo ecology and Geochronology of Hungry Horse Reservoir, Flathead Na-tional Forest, Montana: A First Approximation. Research Report P7. Richland, WA: Applied Paleo science.

———. (1998a). The Central Cascades Paleo ecological Project: Palynological Investigations at Four Sites in the Gifford Pinchot and Mount Hood National Forests, Washington and Oregon. Research Report P9. Richland, WA: Applied Paleo science.

———. (1998b). “Environment.” In Handbook of North American Indians, Vol. 12: Plateau, edited by D. E. Walker, pp. 29–48. Washington, DC: Smithsonian Institution.

———. (2000). “The Discovery and Initial Aanalysis of an Early Holocene Human Skeleton from Kennewick, Washington.” American Antiquity 65:261–316.

———. (2003). Archaeological Data Recovery at 45D0373 and 450K420: Two Okanogan Phase Sites in North Central Washington. Report A- 15. Bothell, WA: Applied Paleo science.

———. (2008). “Technological Change as a Contributor to Craniofacial Differences between Paleo Indians and Later Peoples: A Pacifi c Northwest Example.” Paper presented at the 2008 Visiting Scholar’s Conference, “Physical and Archaeological Variation in the Americas,” Southern Illinois University, Car-bondale, April 2008.

———. (2010a). “Patterns of Death and the Peopling of the Americas.” In Symposio International III: “El Hombre Temprano en America,” edited by J. C. Jiménez López, C. S. Sánchez, A. González González, and F. J. Aguilar Arellano, 53–74. Mexico City: Instituto National de Antropología y Historia: Universidad Nacional Autonóma de México, Instituto de Investigaciones Antropológias: Museo del Desierto, AC. .

———. (2010b). “Peopling the Americas via Multiple Migrations from Beringia: Evi-dence from the Early Holocene of the Columbia Plateau.” In Human Variation in the Americas, edited by B. Auerbach, 51–76. Carbondale: Southern Illinois University Press.

Chatters, J. C., V. L. Butler, M. J. Scott, D. M. Anderson, and D. A. Neitzel. (1995). “A Paleo science Approach to Estimating the Effects of Climatic Warming on Sal-monid Fisheries of the Columbia River Basin.” In Climate Change and Northern Fish Populations, edited by R. J. Beamish, 468–73. Canadian Special Publica-tion of Fisheries and Aquatic Sciences No. 121.

Chatters, J. C., S. Hackenberger, A. Busacca, L. S. Cummings, R. L. Jantz, T. W. Staf-ford, and R. E. Taylor. (2000). “A Second Early Holocene Skull from Eastern Washington, USA?” Current Research in the Pleistocene 17:93–95.

TAM Bousman 13106 complete.indd 59TAM Bousman 13106 complete.indd 59 4/30/12 8:27 AM4/30/12 8:27 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 24: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

60 Chatters, Hackenberger, Prentiss, and Thomas

Chatters, J. C., and D. Pokotylo. (1998). “An Introduction to Plateau Prehis-tory.” In Handbook of North American Indians, Vol. 12: The Plateau, edited by D. E. Walker, 73–80. Washington, DC: Smithsonian Institution.

Chatters, J. C., and A. M. Prentiss. (2010). “Technological and Functional Analysis of Lithic Artifacts.” In Archaeological Data Recovery at the Beech Creek Site (45LE415), Gifford Pinchot National Forest, Washington, edited by C. A. Mack, J. C. Chat-ters, and A. L. Prentiss. Trout Lake, WA: Pacifi c Northwest Region, US Forest Service.

Cheatham, R. (1988). Late Archaic Settlement Patterns in the Long Tom Sub- Basin, Upper Willamette Valley, Oregon. Anthropological Paper No. 39. Eugene: University of Oregon.

Choquette, W. (1996). “Early Postglacial Habitation of the Upper Columbia Region.” In Early Human Occupation in British Columbia, edited by R. L, Carlson and L. Dalla Bona, 45–50. Vancouver: University of British Columbia Press.

Clague, J. J. (1981). “Late Quaternary Geology and Geochronology of British Co-lumbia, Part 2: Summary and Discussion of Radio- Carbon Dated Quaternary History.” Geological Survey of Canada Paper No. 80–35.

Clague, J. J., and R. W. Mathewes. (1989). “Early Holocene Thermal Maximum in Western North America: New Evidence from Castle Peak, British Columbia.” Geology 17:277–80.

Cole, D. L. (1965). “Report on Archaeological Research in the John Day Reservoir Area, 1964.” Eugene: University of Oregon, Museum of Natural History.

Connolly, T. J. (1999). Newberry Crater: A Ten- Thousand- Year Record of Human Occupa-tion and Environmental Change in the Basin- Plateau Borderlands. Anthropological Paper No. 121. Salt Lake City: University of Utah Press.

Connolly. T. J., and P. Barker. (2004). “Basketry Chronology of the Early Holocene in the Northern Great Basin.” In Early and Middle Holocene Archaeology of the Northern Great Basin, edited by D. L. Jenkins, T. J. Connolly, and C. M. Aikens, 241–50. Anthropological Paper No. 62. Eugene: University of Oregon.

Connolly, T. J., and D. L. Jenkins. (1999a). “Great Basin Archaeology of the Early and Middle Holocene in Central Oregon.” Athena Review 2:70–77.

——— and ———. (1999b). “The Paulina Lake Site (35DS34).” In Newberry Crater: A Ten- Thousand- Year Record of Human Occupation and Environmental Change in the Basin- Plateau Borderlands, edited by T. J. Connolly, 86–126. Anthropological Paper No. 121. Salt Lake City: University of Utah Press.

Craven, S. L. (2003). “Lithic Variation in Hafted Bifaces from the Lind Coulee Site (45GR97), Washington.” Unpublished MA thesis, Department of Anthropol-ogy, Washington State University, Pullman.

Cressman, L. S., D. L. Dole, W. A. Davis, T. M. Newman, and D. J. Sheans. (1960). “Cultural Sequences at The Dalles, Oregon: A Contribution to Pacifi c North-west Pprehistory.” Transactions of the American Philosophical Society 100:223–78.

Daughtery, R. D. (1956). “Archaeology of the Lind Coulee Site, Washington.” Proceed-ings of the American Philosophical Society 100:233–78.

———. (1962). “The Intermontane Western Tradition.” American Antiquity 28:144–50.Daugherty, R. D., J. J. Flenniken, and J. M. Welch. (1987a). A Data Recovery Study of

Judd Peak Rockshelters (45- LE- 222) in Lewis County, Washington. Studies in Cul-tural Resource Management No. 8. Portland, OR: USDA Forest Service, Pacifi c Northwest Region.

———, ———, and ———. (1987b). A Data Recovery Study of Layser Cave (45LE223) in Lewis County, Washington. Studies in Cultural Resources Management No. 7. Portland, OR: USDA Forest Service, Pacifi c Northwest Region.

Davis, L. G. (2001). “Lower Salmon River Cultural Chronology: A Revised and Ex-panded Model.” Northwest Anthropological Research Notes 35:229–47.

Davis, L. G., and D. A. Sisson. (1998). “An Early Stemmed Point Cache from the Lower Salmon River Canyon of West- Central Idaho.” Current Research in the Pleistocene 15:12–13.

Davis, S. D. (1989). “Cultural Component 1.” In Hidden Falls Site, Baranof Island, Alaska, edited by S. D. Davis, 31–39. Anchorage: Alaska Anthropological As-sociation.

Dixon, E. J. (2008). “Bifaces from On- Your- Knees Cave, Southeast Alaska.” In Projectile Point Sequences in Northwestern North America, edited by R. L. Carlson and M. P. R. Magne, 11–18. Burnaby, BC: Archeology Press, Simon Fraser University.

Draper, J. A. (1986a). “450K 424 Site Report.” In: The Wells Reservoir Archaeologi-cal Project, Vol. 2: Site Reports, edited by J. C. Chatters. Central Washington

TAM Bousman 13106 complete.indd 60TAM Bousman 13106 complete.indd 60 4/30/12 8:27 AM4/30/12 8:27 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 25: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

61 Pacifi c NorthwestArchaeological Survey, Archaeological Report 86–6. Ellensburg: Central Washington University.

———. “45D0387 Site Report.” (1986b). In The Wells Reservoir Archaeological Project, Vol. 2: Site Reports, edited by J. C. Chatters. Central Washington Archaeologi-cal Survey, Archaeological Report 86–6. Ellensburg: Central Washington University.

Dumond, D. E., and R. Minor. (1983). Archaeology in the John Day Reservoir: The Wildcat Canyon Site. Anthropological Paper No. 30. Eugene: University of Oregon.

Fagan, J. (1988). “Clovis and Western Pluvial Lakes Tradition Lithic Technologies at the Dietz Site, in South Central Oregon.” In Early Human Cccupation in Far Western North America: The Clovis- Archaic Interface, edited by J. Willig, C. M. Ai-kens, and J. L. Fagan, 389–416. Anthropological Paper No. 21. Carson City: Nevada State Museum.

———. (1999). “Analysis of the Lithic Artifact Embedded in the Columbia Park Re-mains.” Report Submitted to the National Park Service, US Department of the Interior, Washington, DC.

Fedje, D. W. (1996). “Early Human Presence in Banff National Park.” In Early Hu-man Occupation in British Columbia, edited by R. L. Carlson and L. Dalla Bona, 33–44. Vancouver: University of British Columbia Press.

Fedje, D. W., A. P. Mackie, D. McLaren, and T. Christensen. (2008). “A Projectile Point Sequence for Haida Gwaii.” In Projectile Point Sequences in Northwestern North America, edited by R. L. Carlson and M. P. R. Magne, 11–18. Burnaby, BC: Archeology Press, Simon Fraser University.

Fedje, D. W., A. P. Mackie, J. B. McSporran, and B. Wilson. (1996). “Early Period Ar-chaeology in Gwaii Haanas: Results of the 1993 Field Program.” In Early Hu-man Occupation in British Columbia, edited by R. L. Carlson and L. Dalla Bona, 133–50. Vancouver: University of British Columbia Press.

Fedje, D. W., M. Magne, and T. Christensen. (2005). “Test Excavations at Raised Beach Sites in Southern Haida Gwaii and Their Signifi cance in Pacifi c Northwest Archaeology.” In Haida Gwaii: Human History and Environment from the Time of the Loon to the Time of the Iron People, edited by D. Fedje and R. Mathewes, 204–44. Vancouver: University of British Columbia Press.

Ferguson, G. J., and W. F. Libby. (1962). “UCLA Radiocarbon Cates.” Radiocarbon 4:109–14.

Galm, J. R., and S. Gough. (2001). “Site 45KT1362, a c. 10,000 yr B.P. Occupation in Central Washington.” Current Research in the Pleistocene 17:29–31.

——— and ———. (2008). “The Projectile Point/Knife Sample from the Sentinel Gap Site.” In Projectile Point Sequences in Northwestern North America, edited by R. L. Carlson and M. P. R. Magne, 209–20. Burnaby, BC: Archaeology Press, Simon Fraser University.

Gehr, K. D. (1980). Late Pleistocene and Recent Archaeology and Geomorphology of the South Sshore of Harney Lake, Oregon. Portland, OR: Portland State University.

Gilbert, T. M. P., D. L. Jenkins, A. Gotherstrom, N. Naveren, J. J. Sanchez, M. Ho-freiter, P. F. Thomsen, J. Binladen, T. F. G. Higham, R. M. Yohe II, R. Parr, L. S. Cummings, and E. Willerslev. (2008). “DNA from pre- Clovis Human Coprolites in Oregon, North America.” Science 320:786–89.

Goebel, T., M. R. Waters, and D. H. O’Rourke. (2008). “The Late Pleistocene Dispersal of Modern Humans in the Americas.” Science 319:1497–1502.

Grabert, G. F. (1968). “North- Central Washington Prehistory: A Final Report on Sal-vage Archaeology in the Wells Reservoir, Part 1.” Reports in Archaeology No. 1. Seattle: University of Washington, Department of Anthropology.

Green, T. J., B. Cochran, T. W. Fenton, J. C. Woods, G. L. Titmus, L. Tieszen, M. A. Davis, and S. J. Miller. (1998). “The Buhl Burial: A Paleo indian Woman from Southern Idaho.” American Antiquity, 63:437–56.

Green, T. J., M. G. Pavesic, J. C. Woods, and G. L. Titmus. (1986). “The DeMoss Burial Locality: Preliminary Observations.” Idaho Archaeologist 9:437–56.

Greene, G. S. (1975). “Prehistoric Utilization of the Channeled Scablands of Eastern Washington.” Unpublished PhD dissertation, Washington State University, Pullman.

Gustafson, C., Gilbow, D., and Daugherty, R. (1979). “The Manis Mastodon Site: Early Man on the Olympic Peninsula.” Canadian Journal of Archaeology 3:157–64.

Gustafson, C. E. (1972). “Faunal Remains from the Marmes Rock Shelter and Related Sites in the Columbia Basin.” PhD dissertation. Ann Arbor, MI: Dissertation Microfi lms.

TAM Bousman 13106 complete.indd 61TAM Bousman 13106 complete.indd 61 4/30/12 8:27 AM4/30/12 8:27 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 26: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

62 Chatters, Hackenberger, Prentiss, and Thomas

Guthrie, R. D. (1982). “Mammals of the Mammoth Steppe as Paleo environmental In-dicators.” In Paleo ecology of Beringia, edited by D. M. Hopkins, J. V. Matthews, C. E. Schweger, and S. B. Young, 307–26. New York: Academic Press.

Hackenberger, S., and R. W. Thompson. (1995). “Archaeological Investigations at the Tryon Creek Site (35WA288), Hells Canyon National Recreation Area (PA 6–92–16–118), Wallowa County, Oregon.” Manuscript on fi le, Wallowa- Whitman National Forest, US Forest Service, Enterprise, OR.

Hanes, R. C. (1988). “Lithic Assemblages of Dirty Shame Rockshelter: Changing Tra-ditions in the Northern Intermontane.” Anthropological Papers 40:206.

Harten, L. B. (1980). “The Osteology of the Human Skeletal Material from the Braden Site, 10- WN- 117, in Western Idaho.” In Anthropological Papers in Memory of Earl H. Swanson, Jr., edited by L. B. Harten, C. N. Warren, and D. R. Tuohy, 130–48. Pocatello: Idaho Museum of Natural History.

Hayden, B., N. Franco, and J. Spafford. (1996). “Lithic Strategies and Design Criteria.” In Stone Tools: Theoretical Insights into Human Prehistory, edited by G. H. Odell, 9–50. New York: Plenum Press.

Hebda, R. J. (1995). “British Columbia Vegetation and Climate History with a Focus on 6 ka BP.” Geographie Physical Quaternaire 49:55–79.

Heinrichs, M. L., R. J. Hebda, I. R. Walker, and S. L. Palmer. (2002). “Postglacial Paleo ecology and Inferred Paleo climate in the Engelmann Spruce–Subalpine Fir Forest of South- Central British Columbia, Canada.” Paleo geography, Paleo-climatology, Paleo ecology 184:347–69.

Hicks, B. A. (2004). “Interpretation.” In Marmes Rockshelter: A Final Report on 11,000 Years of Cultural Use, edited by B. A. Hicks, 389–421. Pullman: Washington State University Press.

Howells, W. W. (1973). “Cranial Variation in Man: A Multivariate Analysis of Patterns of Difference among Human Populations.” Papers of the Peabody Museum of Archaeology and Ethnology, Vol. 79. Cambridge, MA: Harvard University.

Huckleberry, G., B. Lenz, J. Galm, and S. Gough. (2003). “Recent Geoarchaeological Discoveries in Central Washington.” Geological Society of America, Field Guide 4: Western Cordillera and Adjacent Areas, Vol. 4, 237–49.

Huckleberry, G., and J. K. Stein. (1999). “Preliminary Geoarchaeological Stud-ies Associated with Human Remains Found at Columbia Park, Kennewick, Washington.” Report Submitted to the National Park Service, US Department of the Interior, Washington, DC.

Irwin, A. M., and U. Moody. (1978). The Lind Coulee Site (45GR97). Washington Ar-chaeological Research Center Project Report No. 56. Pullman: Washington State University.

Jaehnig, M. E. W. (1984). Archaeological Investigations 45D0273, Chief Joseph Dam Project. Seattle: Offi ce of Public Archaeology, University of Washington.

Jantz, R. L., and D. W. Owsley. (2001). “Variation among Early North American Cra-nia.” American Journal of Physical Anthropology 114:146–55.

Knudson, R. (1983). Organizational Variability in late Paleo - Indian Assemblages. Pullman: Washington State University, Laboratory of Anthropology.

Krantz, G. S. (1979). “Oldest Human Remains from the Marmes Site.” Northwest Anthropological Research Notes 13:159–73.

Lenz, B. R. (in press). “Paleo indian Land Use in the Columbia Plateau.” PhD disserta-tion, University of Leicester, UK.

Leonhardy, F. C. (1970). “Artifact assemblages and archaeological units at Granite Point locality 1 (45WT41). Southeastern Washington. PhD dissertation, Wash-ington State University. Pullman.

Leonhardy, F. C., and D. G. Rice. (1970). “A Proposed Culture Typology for the Lower Snake River, Southeastern Washington.” Northwest Anthropological Research Notes 19:161–68.

Lepofsky, D., and M. Lenert. (2005). “Report on the 2004 Excavations of the MacCallum Site (DhRk2), Aggasiz, b.c. Report on File, Archaeology Banch, Victoria, BC.

Lewarch, D. E., and J. R. Benson. (1989). Archaeological Data Recovery at the Vine Site (45SA85). Evans- Hamilton, Inc., Report for the Mt Adams Ranger District, Gif-ford Pinchot National Forest, USDA- USFS.

Lyman, R. L. (1992). “Infl uences of Mid- Holocene Altithermal climates on mam-malian faunas and human subsistence in Eastern Washington.” Journal of Ethnobiology 12:37–62.

Lyman, R. L., and S. D. Livingston. (1983). “Late Quaternary mammalian zoogeogra-phy of Eastern Washington.” Quaternary Research 20:360–76.

TAM Bousman 13106 complete.indd 62TAM Bousman 13106 complete.indd 62 4/30/12 8:27 AM4/30/12 8:27 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 27: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

63 Pacifi c NorthwestMann, D. H., and T. D. Hamilton. (1995). “Late Pleistocene and Holocene Paleo-environments of the North Pacifi c Coast.” Quaternary Science Reviews 14:449–71.

Matson, R. G. (1996). “The Old Cordilleran Component at the Glenrose Cannery Site.” In Early Human Occupation in British Columbia, edited by R. L, Carlson and L. Dalla Bona, 11–122. Vancouver: University of British Columbia Press.

Matson, R. G., and G. Coupland. (1995). The Prehistory of the Northwest Coast. Orlando, FL: Academic Press.

McClure, R. H. (1992). An Archaeological Assessment of the Beech Creek Site (45LE415), Gifford Pinchot National Forest. Vancouver, WA: USDA Forest Service, Region 6, Gifford Pinchot National Forest.

McPherson, P. J., D. M. Hall, V. J. McGlone, N. J. Nachtwey. (1981). Archaeological Ex-cavation in the Blue Mountains: Mitigation of Sites 35UN52, 35UN74, And 35UN95 in the Vicinity of Ladd Canyon, Union County, Oregon. Boulder, CO: Western Cul-tural Resource Management.

Mehringer, P. J., Jr. (1988). “The Ritchie- Roberts Clovis Cache, East Wenatchee, Washington.” Northwest Science 62:271–72.

Minor, R. (1984). “An Early Complex at the Mouth of the Columbia River.” Northwest Anthropological Research Notes 18:1–22.

Minor, R., and K. A. Toepel. (1986). “Archaeological Assessment of the Bob’s Point Site, (45KL219), Kickitat County, Washington.” Report No. 42. Eugene, OR: Heritage Research Associates.

Miss, C. J., and L. Hudson. (1987). Cultural Resources Collection Analysis: Albeni Falls Project, Northern Idaho. Seattle, WA: US Army Corps of Engineers.

Moss, M. L., and J. M. Erlandson. (1996). “The Pleistocene- Holocene Transition along the Pacifi c Coast of North America.” In Humans at the End of the Ice Age: The Archaeology of the Pleistocene- Holocene Transition, edited by L. G. Strauss, B. V. Er-ickson, J. M. Erlandson, and D. R. Yesner, 272–301. New York: Plenum Press.

Munsell, D. A. (1968). “The Ryegrass Coulee Site.” Unpublished MA thesis, Univer-sity of Washington, Seattle.

Nance, C. R. (1966). “45WT2: An Archaeological Site on the Lower Snake River.” Unpublished MA thesis, Washington State University, Pullman.

Nelson, C. M. (1969). “The Sunset Creek Site (45- KT- 28) and Its Place in Plateau Pre-history.” Report of Investigations No. 47. Pullman: Washington State Univer-sity, Laboratory of Anthropology.

Newman, T. M. (1966). “Cascadia Cave.” Occasional Paper No 18. Pocatello: Idaho State University Museum.

O’Brien, M. J., and R. L. Lyman. (2000). Applying Evolutionary Archaeology: A Systematic Approach. New York: Kluwer/Plenum.

Oetting, A. C. (1994). “Early Holocene Rabbit Drives and Prehistoric Land Use Pat-terns on Buffalo Flat, Christmas Lake Valley, Oregon.” Anthropological Papers 50:155–69.

Pavesic, M. G. (1985). “Cache Blades and Turkey Tails: Piecing Together the Western Idaho Archaic Burial Complex.” In: Stone Tool Analysis: Essays in honor of Don E. Crabtree, edited by M. G. Plew, J. C. Woods, and M. G. Pavesic, 55–89. Albu-querque: University of New Mexico.

Pellatt, M. G., M. J. Smith, R. W. Mathewes, and I. R. Walker. (1998). “Paleo ecology of Postglacial Treeline Shifts in the Northern Cascade Mountains, Canada.” Paleo geography, Paleo climatology, Paleo ecology 141:123–38.

Pennefeather- Obrien, E. E., and M. Strezewski. (2002). “An Initial Description of the Archaeology and Morphology of the Clarks Fork Skeletal Material, Bonner County, Idaho.” North American Archaeologist 23:101–15.

Pettigrew, R. M. (1981). A Prehistoric Culture Sequence in the Portland Basin of the Lower Columbia Valley. Anthropological Paper No. 22. Eugene: University of Oregon.

Pinson, A. O. (2004). “Of Lakeshores and Fry Basin Floors: A Regional Perspective on the Early Holocene Record of Environmental Change and Human Adapta-tion at the Tucker Site.” In Early and Middle Holocene Archaeology of the Northern Great Basin, edited by D. L. Jenkins, C. M. Aikens, and T. J. Connolly, 53–76. Anthropological Paper No. 62. Eugene: University of Oregon.

Powell, J. F. (2004). The First Americans: Race, Evolution, and the Evolution of Native Americans. Cambridge: Cambridge University Press.

Powell, J. F., and J. C. Rose. (1999). “Report on the Osteological Assessment of the ‘Kennewick Man’ Skeleton (CENWW.97. Kennewick).” Report Submitted to the National Park Service, US Department of the Interior, Washington, DC.

Prentiss, A. M., and D. S. Clarke. (2008). “Lithic Technological Organization in an

TAM Bousman 13106 complete.indd 63TAM Bousman 13106 complete.indd 63 4/30/12 8:27 AM4/30/12 8:27 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 28: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

64 Chatters, Hackenberger, Prentiss, and Thomas

Evolutionary Framework: Examples from North America’s Pacifi c Northwest Region.” In Lithic Technology, edited by W. Andrefsky, 257–85. Cambridge: Cambridge University Press.

Randolph, J. E., and M. Dahlstrom. (1977). Archaeological Test Excavations at Bernard Creek Rockshelter. Anthropological Research Manuscript Series No 42, Moscow: University of Idaho.

Reid, K. C., and J. C. Chatters. (1997). “Kirkwood Bar: Passports in Time Excavations at 10IH699 in the Hells Canyon National Recreation Area, Wallowa Whit-man, National Forest.” Project Report No. 28. Pullman, WA: Rain Shadow Research.

Rice, D. G. (1969). Preliminary Report: Marmes Rockshelter Archaeological Site, Southern Columbia Plateau. Pullman: Washington State University Laboratory of An-thropology.

———. (1972). “The Windust Phase in Lower Snake River Region Prehistory. Report of Investigations No. 50. Pullman: Washington State University, Laboratory of Anthropology.

Rice, H. S. (1965). “The Cultural Sequence at Windust Caves.” Unpublished MA thesis, Washington State University, Pullman.

Ridenour, D. I. (2006). “Results of Investigations at We’petes Pa’axat (10NP453), Clearwater River Region, North Central, Idaho.” Unpublished MA thesis, University of Idaho, Moscow.

Rousseau, M. K. (1991). “Landels: An 8500 Year- Old Hunting Camp.” The Midden 23(4):6–9.

Rudolph, T. (1995). The Hetrick Site: 11,000 Years of Prehistory in the Wieser Valley. Boise, ID: Science Applications International Corp.

Salo, L. V. (1987). “Similkameen River Multipurpose Project Feasibility Study Cul-tural Resources Reconnaissance Technical Report.” Seattle, WA: US Army Corps of Engineers.

Sancetta, C., M. Lyle, L. Heusser, R. Zahn, and J. P. Bradbury. (1992). “Late- Glacial to Holocene Changes in Winds, Upwelling, and Seasonal Production of the Northern California Current System.” Quaternary Research 38:359–70.

Sappington, R. L., and S. Schukencht- McDaniel. (2001). “Wewukiyepuh (10- NP- 336): Contributions of an Early Holocene Windust Phase Site to Lower Snake River Prehistory.” North American Archaeologist 22:353–70.

Sargeant, K. E. (1973). “The Haskett Tradition: A View from Redfi sh Overhang.” MA thesis in Anthropology, Idaho State University, Pocatello.

Schalk, R. F., ed. (1995). Archaeological Investigations PGT- PG&E Pipeline Expansion Proj-ect, Idaho, Washington, Oregon, and California, Vol. 4: Synthesis of Findings. Fresno, CA: INFOTEC Research, Inc.

Sheppard, J. C., P. Wigand, C. E. Gustafson, and M. Rubin. (1987). “A Re- Evaluation of the Marmes Rockshelter Radiocarbon Chronology. American Antiquity 52:118–25.

Shiner, J. L. (1961). “The McNary Reservoir: A Study in Plateau Srchaeology.” Bu-reau of American Ethnology Bulletin No. 179, 149–266. Washington, DC: Smithsonian Institution.

Sprague, R., and J. D. Combes. (1966). “Excavations in the Little Goose and Lower Granite Dam Reservoirs, 1965.” Report of Investigations No. 37. Pullman: Washington State University, Laboratory of Anthropology.

Stryd, A., and M. Rousseau. (1996). “Early Prehistory of the Fraser- Thompson River Area.” In Early human occupation in British Columbia, edited by R. L. Carlson and L. Dalla Bona, 177–204. Vancouver: University of British Columbia Press.

Swanson, E. H., Jr., and P. G. Sneed. (1966). “Birch Creek Papers No. 3: The Archae-ology of the Shoup Rock Shelters in East Central Idaho.” Occasional Paper No. 17. Pocatello: Idaho State University Museum.

Tamm, E., T. Kivisild, M. Reidla, M. Metspalu, D. G. Smith, C. J. Mulligan, C. M. Bravi, O. Richards, C. Martinez- Labarga, E. K. Khusnutdinova, S. A. Federova, M. V. Golubenko, V. A. Stepanov, M. A. Gubina, S. I. Zha-danov, L. P. Ossipova, L. Damba, M. I. Voevoda, J. E. Dipierri, R. Villems, and R. S. Malhi. (2007). “Beringian Standstill and Spread of Native American Founders.” PLos ONE 2, e829.

Taylor, E. R., D. Kirner, J. Southon, and J. Chatters. (1998). “Radiocarbon Dates of Kennewick Man.” Science 280:1171–72.

Thoms, A. V. (1989). “Northern Roots of Hunter- Gatherer Intensifi cation: Camas and the Pacifi c Northwest.” PhD Dissertation. Ann Arbor, MI: University Micro-fi lms.

TAM Bousman 13106 complete.indd 64TAM Bousman 13106 complete.indd 64 4/30/12 8:27 AM4/30/12 8:27 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution

Page 29: The Paleo indian to Archaic Transition in the Pacifi c Northwest … · 2013. 2. 13. · Stone implements include large, thin, broad- bladed stemmed and lanceolate pro-jectile points,

65 Pacifi c NorthwestTorrence, R. (1983). “Time Budgeting and Hunter- Gatherer Technology.” In Hunter- Gatherer Economy on Prehistory: A European Perspective, edited by G. Bailey, 11–22. Cambridge: Cambridge University Press.

Valley, D. R. (1975). Excavations at the Rock Island Overlook Site. Pullman: Washington State University, Laboratory of Anthropology.

Waitt, R. B., Jr., and R. M. Thorsen. (1983). “The Cordilleran Ice Sheet in Washing-ton, Idaho, and Montana.” In: Late Quaternary Environments of the United States, edited by S. C. Porter and H. E. Wright, Vol. 1, 53–70. Minneapolis: Univer-sity of Minnesota Press.

Wallthal, J. A. (1999). “Mortuary Behavior and Early Holocene Land Use in the North American Midcontinent.” North American Archaeologist 20:1–30.

Warren, C. N., A. L. Bryan, and D. R. Tuohy. (1963). “The Goldendale Site and Its Place in Plateau Prehistory.” Tebiwa 32:68–185.

Weisz, G. J. (2006). “Bola Stones and Line Weights from the Hornby Creek Site, Northern Idaho.” Idaho Archaeologist 29:19–24.

Wilson, Z. W. (2008). “An Analysis of the Lind Coulee (45GR97) Faunal Assem-blage.” Unpublished MA Thesis, Washington State University, Pullman.

TAM Bousman 13106 complete.indd 65TAM Bousman 13106 complete.indd 65 4/30/12 8:27 AM4/30/12 8:27 AM

Not for reproduction or distribution Not for reproduction or distribution

Not for reproduction or distribution

Not for reproduction or distribution Not for reproduction or distribution