Biface Traditions of Northern Alaska and Their Role in the Peopling of the Americas

19
105 Chapter 6 Biface Traditions of Northern Alaska and Their Role in the Peopling of the Americas Heather L. Smith 1 , Jeffrey T. Rasic 2 , and Ted Goebel 1 ABSTRACT In northeastern Beringia, the area today made up of northern Alaska (USA) and Yukon (Canada), archaeolo- gists typically refer to the earliest archaeological assemblages dominated by bifacial technology as “Northern Paleoindian.” These assemblages mostly date to during and immediately after the Younger Dryas cold period, 12,900–11,200 calendar years ago, although at least one and possibly three occupations pre-date 13,000 calendar years ago and rival Clovis in age. Distinctive lanceolate bifacial points dominate these assemblages; variation in size, shape, production technique, and basal treatment of these points has led archaeologists to define three assemblage groups, the fluted-point, Mesa, and Sluiceway complexes. These projectile points were components of technological systems that emphasized high-quality lithic raw materials used in a formal bifacial industry. Scarce faunal remains and other evidence suggest a subsistence regime centered on the hunt- ing of herd animals such as caribou and bison, and land-use patterns characterized by seasonally structured movements and high logistical mobility. Standardized wedge-shaped core and microblade technology has not been unequivocally tied to the Northern Paleoindian complexes, giving them a character different from most other early Beringian sites and more strongly suggesting a link to late-Pleistocene technocomplexes in the midcontinent. However, careful technological and morphometric studies are needed to assess the relationship of these early northern Beringian industries to those of like age in the temperate latitudes of North America. KEYWORDS: Beringia, Fluted points, Mesa complex, Sluiceway complex 1 Center for the Study of the First Americans, Department of Anthropology, Texas A&M University, 4352-TAMU, College Station, TX 77843. 2 National Park Service, 4175 Geist Road, Fairbanks, AK 99709. Corresponding author e-mail: 1 [email protected] its southern and northern foothills, and Arctic Coastal Plain. Today this region is largely devoid of trees and covered by an open tundra landscape; during the late Pleistocene/early Ho- locene, paleoecological evidence suggests a steppe-tundra existed, home to large fauna including bison, caribou, horse, Dall sheep, and musk oxen (Mann et al. 2001, 2013; Rasic and Matheus 2007). Northern Alaska and Yukon are virtually road-less; therefore expense and logistics have hampered ar- chaeological research. In contrast to the deep loess profiles common in central Alaska, sites are typically in shallow or surface contexts notoriously difficult to date and lacking in subsistence remains. Until just two decades ago, no archaeo- logical sites unequivocally predating 12,000 calendar years Introduction The goal of this paper is to bring readers up to date on the terminal-Pleistocene/earliest-Holocene archaeological record of arctic eastern Beringia, namely northern Alaska, U.S.A., and northern Yukon, Canada, from about 100 km below the Arctic Circle north to the Arctic Ocean. As such, the study area encompasses the Seward Peninsula, Brooks Range and Biface Traditions of Northern Alaska and Their Role in the Peopling of the Americas

Transcript of Biface Traditions of Northern Alaska and Their Role in the Peopling of the Americas

105

Chapter 6

Biface Traditions of Northern Alaska and Their Role in the Peopling

of the Americas Heather L. Smith1, Jeffrey T. Rasic2, and Ted Goebel1

ABsTRAcTIn northeastern Beringia, the area today made up of northern Alaska (USA) and Yukon (Canada), archaeolo-gists typically refer to the earliest archaeological assemblages dominated by bifacial technology as “Northern Paleoindian.” These assemblages mostly date to during and immediately after the Younger Dryas cold period, 12,900–11,200 calendar years ago, although at least one and possibly three occupations pre-date 13,000 calendar years ago and rival Clovis in age. Distinctive lanceolate bifacial points dominate these assemblages; variation in size, shape, production technique, and basal treatment of these points has led archaeologists to define three assemblage groups, the fluted-point, Mesa, and Sluiceway complexes. These projectile points were components of technological systems that emphasized high-quality lithic raw materials used in a formal bifacial industry. Scarce faunal remains and other evidence suggest a subsistence regime centered on the hunt-ing of herd animals such as caribou and bison, and land-use patterns characterized by seasonally structured movements and high logistical mobility. Standardized wedge-shaped core and microblade technology has not been unequivocally tied to the Northern Paleoindian complexes, giving them a character different from most other early Beringian sites and more strongly suggesting a link to late-Pleistocene technocomplexes in the midcontinent. However, careful technological and morphometric studies are needed to assess the relationship of these early northern Beringian industries to those of like age in the temperate latitudes of North America.

KEYWORDs: Beringia, Fluted points, Mesa complex, Sluiceway complex

1 Center for the Study of the First Americans, Department of Anthropology, Texas A&M University, 4352-TAMU, College Station, TX 77843.

2 National Park Service, 4175 Geist Road, Fairbanks, AK 99709. Corresponding author e-mail: [email protected]

its southern and northern foothills, and Arctic Coastal Plain. Today this region is largely devoid of trees and covered by an open tundra landscape; during the late Pleistocene/early Ho-locene, paleoecological evidence suggests a steppe-tundra existed, home to large fauna including bison, caribou, horse, Dall sheep, and musk oxen (Mann et al. 2001, 2013; Rasic and Matheus 2007). Northern Alaska and Yukon are virtually road-less; therefore expense and logistics have hampered ar-chaeological research. In contrast to the deep loess profiles common in central Alaska, sites are typically in shallow or surface contexts notoriously difficult to date and lacking in subsistence remains. Until just two decades ago, no archaeo-logical sites unequivocally predating 12,000 calendar years

IntroductionThe goal of this paper is to bring readers up to date on the terminal-Pleistocene/earliest-Holocene archaeological record of arctic eastern Beringia, namely northern Alaska, U.S.A., and northern Yukon, Canada, from about 100 km below the Arctic Circle north to the Arctic Ocean. As such, the study area encompasses the Seward Peninsula, Brooks Range and

Biface Traditions of Northern Alaska and Their Role in the Peopling of the Americas

106

ago (cal yr BP) were known from the area, but since then at least 11 sites have been found, excavated, and dated, provid-ing us with our first glimpses into the lives of arctic eastern Beringia’s late Ice Age humans (Figure 6.1). Archaeologists working in the region have described three lithic complexes from the terminal Pleistocene/early Holocene record: a Northern fluted-point complex, the Sluiceway complex, and the Mesa complex. Each includes a distinctive lanceolate bifacial projectile point form, and associated radiocarbon ages range from approximately 13,200 to 10,000 cal yr BP (all dates in this paper were cali-brated with Oxcal 4.2). Often these complexes have been subsumed within a broader Paleoindian tradition that also encompasses, and thereby suggests historic links to, Paleo-indian complexes in midcontinental North America such as Clovis, Agate Basin, and Hell Gap. The similarities between

among Alaskan archaeologists, but emphasize its reference to a time period and basic lithic technologies. A considerable amount of discussion has surrounded the question of how the “Northern Paleoindian” complexes, which occupied the northern portion of the sole terrestrial connection between Asia and America, related to the early assemblages in midcontinental North America. Do they pre-date Paleoindians of temperate North America and represent ancestral founding populations (e.g., Clark and Clark 1983; Kunz and Reanier 1994), or do they postdate the Pleisto-cene midcontinent occupations and represent a migration of northern Plains bison hunters into the Arctic or northward diffusion of Paleoindian bifacial-point weaponry (e.g., Clark 1984; Hoffecker 2011; Hoffecker and Elias 2007), or do they represent adaptive convergence due to comparable hunting of gregarious ungulates (Meltzer 1993)?

Figure 6.1 Map showing locations of archaeological sites mentioned in text.

north and south that provide the foundation for this argu-ment include not only the technology used to manufacture projectile points, but also other aspects of stone toolkits, technological organization, and possible parallels in eco-nomic and subsistence strategies (Bever 2000, 2001; Hof-fecker 2011; Kunz et al. 2003; Loy and Dixon 1998). The term Paleoindian is problematic in the Far North for several rea-sons, however. The term “Indian” connotes specific ethnic affinities that remain unfounded for these early archaeolog-ical assemblages. Use of the term across this tremendous geographic range in the absence of clearly dated sites in western Canada implies a degree of cultural connection that has yet to be demonstrated. Despite these limitations we use the moniker here in deference to its common use

We organize the paper into the following sections. First, we briefly present the Northern Paleoindian sites that have been radiocarbon dated to at least 12,000 cal yr BP, reviewing their locations, contexts, chronologies, and cultural remains. We then synthesize this information, developing a regional chronology, assessing projectile-point variability, and inter-preting technological, subsistence and settlement organiza-tion of the earliest inhabitants of arctic eastern Beringia.

The sites

Serpentine Hot SpringsThe Serpentine fluted-point site (BEN-192) is located in northwest Alaska on the Seward Peninsula ~150 km north

Smith et al.

107

of Nome. Results of testing and excavation are described in Young and Gilbert-Young (2007) and Goebel et al. (2013). The site contains a buried cultural component in a stratigraphic context, and dates on charcoal from features tentatively in-terpreted as hearths range from 12,400 to 9900 cal yr BP (Table 6.1). Dates just on willow charcoal (instead of shrub birch or Ericaceae) suggest an age of 12,400–12,000 cal yr BP. Among the discoveries in the excavation are four fragments of fluted points. Technological activities focused on finishing and refurbishing bifacial points; numerous “channel” flakes indicate that final fluting occurred on site. Associated faunal remains are carbonized and poorly preserved; identifications include small artiodactyl (e.g., caribou) and hare-sized mam-mal. Serpentine is considered to represent a hunting lookout occupied for a short time, perhaps on multiple occasions.

Raven Bluff Raven Bluff (DEL-402) is located in northwestern Alaska, ap-proximately 160 km north of Kotzebue. Analyses of archaeo-logical materials are ongoing, but a basic outline of the site chronology and assemblage contents are known and pre-sented in Hedman (2010). Of relevance to this discussion is the discovery of a fluted projectile point in a relatively deeply bur-ied, stratified context. The lower of two strata contained the fluted point, manufacturing rejects from initial stages of biface production, abundant flaking debris, and faunal remains of pri-marily caribou. Ten AMS radiocarbon dates on bone collagen, eight of which are firmly associated with the lower stratum, bracket the fluted point between 9800 ± 60 (11,342–11,102 cal yr BP) and 10,720 ± 50 (12,131–11,630 cal yr BP) 14C yr BP (Table 6.1). Good-quality chert dominates the lithic assemblage and is abundant in the DeLong Mountains that surround Raven Bluff as well as in the creek bottom adjacent to the site.

Tuluaq HillThe Tuluaq Hill site (DEL-360) is located along Wrench Creek in the Noatak River basin, northwest Alaska (Rasic 2008). The surrounding landscape includes a series of bedrock chert quarries which include a black chert that dominates the lithic assemblage (Malyk-Selivanova et al. 1998; Rasic 2011). Al-though relatively shallow, 50 cm of frost-heaved stratigraphy was sufficiently intact to have preserved a hearth feature in close association with distinctive Sluiceway projectile points and > 400 biface manufacturing rejects (Table 6.2). Willow charcoal directly from the hearth yielded four AMS radiocar-bon dates ranging from 11,200 ± 40 to 11,120 ± 40 14C yr BP (13,250–12,801 cal yr BP), indicating the age of the pri-mary site occupation. Tuluaq Hill was a stone-tool workshop; weapon repair was also an important activity, as evidenced by 65 broken and damaged Sluiceway points.

NR-5 The NR-5 site (XBM-017) is located near the confluence of the Noatak and Kugururok rivers in northwest Alaska. Accord-ing to Anderson (1972) and Rasic (2008), the site is situated on a hill that rises 70 m above surrounding lakes and low-

lands. Twenty-four bifacial tools, 11 of which are Sluiceway-type projectile points or point fragments, flake tools, and minimal late-stage bifacial flaking debris, were buried 30–60 cm below surface in a deposit of silt and sand. Microblade midsections were found at similar depths as the diagnostic Sluiceway artifacts, but not in direct association with them. Geological ages on the strata and approximate ages for depo-sitional episodes were obtained from thinly and widely dis-persed charcoal presumably from wildfires. These provided four radiocarbon samples yielding AMS dates from 12,020 to 9421 cal yr BP (Table 6.1). An AMS date of 4330 ± 40 14C yr BP on charcoal from a rodent-disturbed area was dismissed, while an AMS date of 8450 ± 40 14C yr BP was from above the artifact-bearing stratum. Raw-material types represented in the lithic assemblage include a variety of cherts, obsidian from the Batza Tena obsidian source, and a small amount of basalt (Rasic 2008). Activities indicated for the NR-5 site sug-gest it served as a briefly occupied hunting overlook.

Nat PassThe Nat Pass site (MIS-495) is located just south of the Brooks Range in northwest Alaska, 12 km east of the Nimiuktuk- Anisak River divide. It rests on a flat bench of a 40-m-tall hill providing a sweeping view of the Brooks Range to the north and Anisak valley to the south (Rasic 2008). Artifacts include three Sluiceway projectile-point fragments, bifacial blanks and preforms, and an oval-platform microblade core preform, all from a surface/near-surface context. Deposition on top of the exposed hilltop was sufficient to bury and preserve char-coal, a few small (non-diagnostic) burned bone fragments, and several burned flakes, together suggesting the former presence of a hearth. Two pieces of charcoal were submitted for AMS analysis, yielding ages of 9910 ± 40 (11,595–11,221 cal yr BP) and 10,010 ± 40 14C yr BP (11,709–11,287 cal yr BP) (Rasic 2008). Although the charcoal and burned bone in-dicate a late-Pleistocene/early-Holocene use of this location, the broad scatter and overlapping clusters of artifacts on the surface of such a prominent landform suggest this location was occupied repeatedly, and the precise nature of associa-tions among artifacts and between artifacts and radiocarbon dates is uncertain.

Irwin SluicewayThe Irwin Sluiceway site (XHP-496), located along the Anisak River in the Noatak National Preserve, northwest Alaska, is reported in Rasic (2008) and discussed in Saleeby (2010) and Rasic (2011). The site is situated on the southeastern edge of a rise 70 m above Buccaneer Creek that provides an expan-sive view of the surrounding landscape. Artifacts are from a single hearth-centered scatter, in a shallowly buried context roughly 15–20 cm below surface. Two AMS 14C dates on char-coal from a clearly delineated hearth feature provided ages of 10,050 ± 70 and 10,060 ± 80 14C yr BP (11,970-11,282 cal yr BP) (Table 6.1), while a third AMS 14C date of 9550 ± 50 (11,101–10,705 cal yr BP) could represent contamination by younger carbon (Rasic 2008). The 17-m2 excavation yielded

Biface Traditions of Northern Alaska and Their Role in the Peopling of the Americas

108

Table 6.1 Radiocarbon and calibrated dates of Paleoindian components from sites mentioned in text.

site Age, 14c yr BP Age, cal yr BP (1σ) Age, cal yr BP (2σ) sample # Material Reference

Raven Bluff 10,130 ± 70 11,971–11,619 12,032–11,403 UCI-67492 AMS bone collagen Hedman 2010Raven Bluff 10,530 ± 80 12,589–12,390 12,624–12,141 UCI-67490 AMS bone collagen Hedman 2010Raven Bluff 10,210 ± 60 12,046–11,816 12,131–11,630 Beta-248992 AMS bone collagen Hedman 2010Raven Bluff 10,120 ± 70 11,969–11,508 12,024–11,398 Beta-266923 AMS bone collagen Hedman 2010Raven Bluff 10,720 ± 50 12,667–12,580 12,731–12,555 Beta-266925 AMS bone collagen Hedman 2010Raven Bluff 10,250 ± 70 12,112–11,824 12,381–11,718 UCI-67489 AMS bone collagen Hedman 2010Raven Bluff 10,280 ± 80 12,375–11,829 12,406–11,756 UCI-67488 AMS bone collagen Hedman 2010Raven Bluff 10,500 ± 70 12,571–12,228 12,597–12,140 UCI-67491 AMS bone collagen Hedman 2010Serpentine Hot Springs 10,295 ± 25 12,120–12,030 12,363–11,973 UCIAMS-90959 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 10,290 ± 25 12,109–12,010 12,210–11,971 UCIAMS-90960 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 10,370 ± 25 12,379–12,132 12,387–12,092 UCIAMS-90961 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 10,320 ± 25 12,210–12,045 12,377–11,998 UCIAMS-90962 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 10,275 ± 25 12,083–11,998 12,135–11,845 UCIAMS-90963 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 9940 ± 25 11,385–11,267 11,591–11,247 UCIAMS-77097 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 9375 ± 25 10,655–10,567 10,680–10,519 UCIAMS-77098 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 10,270 ± 25 12,078–11,993 12,130–11,841 UCIAMS-90953 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 9985 ± 25 11,598–11,321 11,608–11,285 UCIAMS-90954 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 9900 ± 25 11,603–11,398 11,695–11,330 UCIAMS-90955 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 10,015 ± 25 10,147–9914 10,164–9901 UCIAMS-90956 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 8875 ± 25 11,341–11,261 11,394–11,250 UCIAMS-909571 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 9930 ± 25 12,379–12,142 12,391–12,095 UCIAMS-90958 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 10,375 ± 30 12,372–12,046 12,381–12,004 UCIAMS-90965 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 10,325 ± 35 11,763–11,408 11,962–11,357 UCIAMS-90966 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 10,075 ± 50 12,521–12,169 12,531–12,128 UCIAMS-106037 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 10,440 ± 40 10,658–10,524 10,693–10,508 UCIAMS-106038 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 9370 ± 35 10,587–10,501 10,662–10,427 UCIAMS-106036 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 9335 ± 35 10,651–10,520 10,689–10,499 UCIAMS-106036 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 9360 ± 35 10,645–10,510 10,675–10,437 UCIAMS-106040 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 8920 ± 35 10,178–9937 10,190–9914 UCIAMS-106041 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 9345 ± 35 11,325–11,256 11,391–11,246 UCIAMS-106042 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 9920 ± 25 11,396–11,285 11,601–11,261 UCIAMS-77095 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 9960 ± 25 11,305–11,247 11,355–11,236 UCIAMS-77096 AMS charcoal Goebel et al. 2013Serpentine Hot Springs 10,060 ± 40 11,742–11,408 11,813–11,353 Beta-208367 AMS charcoal Young and Gilbert- Young 2007Serpentine Hot Springs 10,250 ± 60 12,105–11,826 12,376–11,754 Beta-208368 AMS charcoal Young and Gilbert- Young 2007Serpentine Hot Springs 9480 ± 40 11,053–10,607 11,068–10,586 Beta-208369 AMS charcoal Young and Gilbert- Young 2007Serpentine Hot Springs 10,250 ± 60 12105–11826 12,376–11,754 Beta-208370 AMS charcoal Young and Gilbert- Young 2007Putu 5700 ± 190 6718–6300 6985–6020 GaK-4941 Charcoal Alexander 1987Putu 6090 ± 430 7418–6505 7823–6004 GaK-4939 Soil Alexander 1987Putu 8454 ± 130 9549–9288 9733–9034 WSU-1318 Soil Alexander 1987Putu 11,470 ± 500 13,886–12,751 15,078–12,394 SI-2382 Charcoal Alexander 1987Putu 8810 ± 60 10,117–9700 10,158–9631 Beta-69901, AMS charcoal Reanier 1995 CAMS-11038Bedwell 10,490 ± 70 12,567–12,223 12,590–12,136 Beta-69895, AMS charcoal Reanier 1995 CAMS-11032Hilltop 10,360 ± 60 12,381–12,095 12,513–11,994 Beta-69897, AMS charcoal Reanier 1995 CAMS-11034Mesa East Ridge 9740 ± 50 11,226–11,147 11,247–10,884 Beta-120400 AMS charcoal Kunz et al. 2003Mesa East Ridge 10,030 ± 40 11,684–11,398 11,750–11,329 Beta-125998 AMS charcoal Kunz et al. 2003Mesa East Ridge 10,080 ± 40 11,765–11,410 11,956–11,397 Beta-125997 AMS charcoal Kunz et al. 2003Mesa East Ridge 9780 ± 40 11,234–11,192 11,254–11,160 Beta-130577 AMS charcoal Kunz et al. 2003Mesa East Ridge 9930 ± 40 11,390–11,256 11,601–11,236 GX-26567-AMS AMS charcoal Kunz et al. 2003Mesa Saddle 9730 ± 80 11,240–10,885 11,269–10,779 Beta-36805 AMS charcoal Kunz et al. 2003Mesa Saddle 9945 ± 75 11,601–11,246 11,710–11,220 Beta-50430 AMS charcoal Kunz and Reanier 1995

Smith et al.

109

Mesa Saddle 9990 ± 80 11,613–11,286 11,810–11,238 Beta-55282 AMS charcoal Kunz and Reanier 1995Mesa Saddle 10,070 ± 60 11,761–11,405 11,963–11,336 Beta-69898 AMS charcoal Kunz and Reanier 1995Mesa Saddle 9860 ± 50 11,305–11,215 11,393–11,197 Beta-120399 AMS charcoal Kunz et al. 2003Mesa Saddle 9920 ± 50 11,391–11,246 11,604–11,225 Beta-120398 AMS charcoal Kunz et al. 2003Mesa Saddle 9800 ± 60 11,255–11,180 11,342–11,102 Beta-120793 AMS charcoal Kunz et al. 2003Mesa Saddle 9950 ± 60 11,598–11,250 11,696–11,234 Beta-133354 AMS charcoal Kunz et al. 2003Mesa Saddle 10,180 ± 60 11,996–11,759 12,089–11,509 Beta-133353 AMS charcoal Kunz et al. 2003Mesa Saddle 10,080 ± 50 11,768–11,408 11,965–11,367 Beta-142261 AMS charcoal Kunz et al. 2003Mesa A 10,090 ± 85 11,951–11,405 12,009–11,315 Beta-50428 AMS charcoal Kunz and Reanier 1995Mesa A 10,080 ± 50 11,768–11,408 11,965–11,367 Beta-84650 AMS charcoal Kunz et al. 2003Mesa A 10,230 ± 60 12,071–11,824 12,371–11,649 Beta-95600 AMS charcoal Kunz et al. 2003Mesa A 10,260 ± 110 12,374–11,767 12,522–11,412 Beta-96070 AMS charcoal Kunz et al. 2003Mesa A 10,150 ± 130 12,050–11,410 12,376–11,271 Beta-96069 AMS charcoal Kunz et al. 2003Mesa A 10,080 ± 120 11,954–11,396 12,062–11,251 Beta-96068 AMS charcoal Kunz et al. 2003Mesa A 9850 ± 150 11,616–11,106 11,954–10,773 Beta-96067 AMS charcoal Kunz et al. 2003Mesa A 10,090 ± 110 11,954–11,403 12,049–11,268 Beta-96066 AMS charcoal Kunz et al. 2003Mesa A 9810 ± 110 11,399–10,898 11,700–10,786 Beta-96065 AMS charcoal Kunz et al. 2003Mesa A 10130 ± 60 11,966–11,625 12,028–11,405 Beta-95914 AMS charcoal Kunz et al. 2003Mesa A 10,080 ± 60 11,808–11,406 11,970–11,348 Beta-95913 AMS charcoal Kunz et al. 2003Mesa A 10,130 ± 50 11,963–11,641 12,014–11,408 Beta-118585 AMS charcoal Kunz et al. 2003Mesa A 10,040 ± 50 11,699–11,400 11,807–11,312 Beta-118584 AMS charcoal Kunz et al. 2003Mesa A 10,050 ± 50 11,706–11,404 11,819–11,319 Beta-118583 AMS charcoal Kunz et al. 2003Mesa A 10,100 ± 50 11,953–11,417 11,970–11,403 Beta-118582 AMS charcoal Kunz et al. 2003Mesa A 10,170 ± 50 11,974–11,765 12,053–11,623 Beta-118581 AMS charcoal Kunz et al. 2003Mesa A 10,000 ± 50 11,606–11,339 11,709–11,267 Beta-119100 AMS charcoal Kunz et al. 2003Mesa A 10,120 ± 50 11,957–11,623 11,987–11,406 Beta-142262 AMS charcoal Kunz et al. 2003Mesa B 10,060 ± 70 11,753–11,404 11,967–11,310 Beta-52606 AMS charcoal Kunz and Reanier 1995Mesa B 11,660 ± 80 13,619–13,400 13,728–13,326 Beta-55286 AMS charcoal Kunz and Reanier 1995Mesa B 10,000 ± 80 11,692–11,309 11,821–11,240 Beta-55285 AMS charcoal Kunz and Reanier 1995Mesa B 9930 ± 80 11,600–11,237 11,709–11,207 Beta-55284 AMS charcoal Kunz and Reanier 1995Mesa B 10,240 ± 80 12,114–11,771 12,385–11,627 Beta-55283 AMS charcoal Kunz and Reanier 1995Mesa B 11,190 ± 70 13,180–12,962 13,279–12,861 Beta-57430 AMS charcoal Kunz and Reanier 1995Mesa B 9900 ± 70 11,400–11,219 11,613–11,200 Beta-57429 AMS charcoal Kunz and Reanier 1995Mesa B 10,050 ± 90 11,762–11,357 11,971–11,266 Beta-69900 AMS charcoal Kunz and Reanier 1995Mesa B 9900 ± 80 11,593–11,214 11,701–11,189 Beta-69899 AMS charcoal Kunz and Reanier 1995Mesa B 9980 ± 60 11,603–11,286 11,709–11,248 Beta-84649 AMS charcoal Kunz et al. 2003Engigstciak 9870 ± 180 11,742–11,106 11,989–10,755 RIDDL-362 AMS bone collagen Cinq-Mars et al. 1991Engigstciak 9770 ± 180 11,399–10,779 11,822–10,592 RIDDL-281 AMS bone collagen Cinq-Mars et al. 1991Engigstciak 9400 ± 230 11,075–10,300 11,305–9961 RIDDL-319 AMS bone collagen Cinq-Mars et al. 1991Tuluaq Hill 11,110 ± 80 13,115–12,895 13,182–12,729 Beta-122323 AMS charcoal Rasic and Gal 2000Tuluaq Hill 11,120 ± 40 13,100–12,946 13,142–12,801 Beta-159913 AMS charcoal Rasic 2011Tuluaq Hill 11,180 ± 80 13,173–12,946 13,276–12,810 Beta-122322 AMS charcoal Rasic and Gal 2000Tuluaq Hill 11200 ± 40 13,192–12,979 13,250–12,911 Beta-133393 AMS charcoal Rasic and Gal 2000Tuluaq Hill 11,160 ± 40 13,128–12,962 13,202–12,865 Beta-159915 AMS charcoal Rasic 2011Tuluaq Hill 11,200 ± 40 13,192–12,979 13,250–12,911 Beta-159914 AMS charcoal Rasic 2011Irwin Sluiceway 9550 ± 50 11,071–10,751 11,101–10,705 Beta-120696 AMS charcoal Rasic 2011Irwin Sluiceway 10,050 ± 70 11,748–11,398 11,958–11,285 Beta-134677 AMS charcoal Rasic 2011Irwin Sluiceway 10,060 ± 80 11,765–11,395 11,970–11,282 Beta-131336 AMS charcoal Rasic 2011NR-52 9640 ± 300 11,403–10,502 12,020–10,234 Beta-146116 AMS charcoal Rasic 2011NR-52 9570 ± 60 11,080–10,764 11,144–10,710 Beta-146117 AMS charcoal Rasic 2011Nat Pass 9910 ± 40 11,337–11,244 11,595–11,221 Beta-165298 AMS charcoal Rasic 2011Nat Pass 10,010 ± 40 11,610–11,359 11,709–11,287 Beta-165299 AMS charcoal Rasic 2011Nogahabara I 10,780 ± 70 12,724–12,594 12,854–12,565 CAMS-107314 AMS bone collagen Odess and Rasic 2007Nogahabara I 11,815 ± 40 13,766–13,601 13,800–13,475 CAMS-112709 AMS bone collagen Odess and Rasic 2007

1Oxcal v4.2.2 Bronk Ramsey (2013); r:5 Atmospheric data from Reimer et al. (2009).2NR-5 dates are geological dates only and roughly provide limiting ages for the Sluiceway component.

Table 6.1 Cont’d.

site Age, 14c yr BP Age, cal yr BP (1σ) Age, cal yr BP (2σ) sample # Material Reference

Biface Traditions of Northern Alaska and Their Role in the Peopling of the Americas

110

Geomorphic site setting Depositional setting Lithic assemblage Faunal remains References

Table 6.2 Qualitative data on sites highlighted in text.

Serpentine Hot Springs

Overlook

Overlook

Overlook

Overlook

Overlook

Quaternary deposits reached 70 cm thick and consisted of three natural strata overlying Cretaceous granite. Unit 1, at the base of the profile, was a colluvial deposit of grussy, clayey silt representing a mixture of reworked, wind-blown silts and sands as well as gruss particles from degrading bedrock. Above this, Unit 2 was a clayey silt to silt package of aeolian loess, minimally affected by colluvial processes. Unit 3 was a grussy sandy silt, colluvial in ori-gin, and mantled by an O horizon. Cryogenic features resulting from ice-lens formation and solifluction, as well as occasional rodent-burrow casts, were present in all three units; however, the strata were not obviously inter-mixed and stratigraphic integrity between units remained intact. The main cultural component was in Unit 2, but a few artifacts also occurred near the top of Unit 3, typically at its contact with the O horizon.

5 bifacial-point fragments, 4 fluted-point basal fragments (from excavation), 3 fluted-point basal fragments (from surface), 15 biface fragments, 7 flake tools, channel flakes, spurred gravers, 3 bladelet fragments, bladelets, microblade fragments, 1 platform-rejuvenation tablet from a conical bladelet core, 1,481 pieces of debitage

Ungulate, small mammal

Goebel et al. 2013; Young and Gilbert-Young 2007

Raven Bluff Stratified sediments up to 160 cm deep were preserved within strata that had been subject to only minimal post-depositional disturbance from cryoturbation. Cultural material occurred up to 120 cm below the surface, terminating just above a culturally sterile gray clay stratum.

1 fluted-point preform, chan-nel flakes, bifacial flaking debris, blade cores, several medial micro blade fragments

Caribou, small mammal

Hedman 2010

Tuluaq Hill Stratum IV, a surface lag deposit, was the up-permost unit and consisted of an extremely gravelly sandy loam up to 25 cm thick and contained abundant artifacts. Stratum III was a dark yellowish brown sandy loam of eolian origin measuring 1–25 cm thick and contained common artifacts. Stratum II was a dark brown silty loam 1–30 cm thick, which also contained artifacts. Stratum I was a gravelly sandy loam, usually 20–35 cm thick, with clasts reaching cobble size and incorporated into shale/lime-stone regolith and devoid of artifacts.

65 Sluiceway point/fragments, 401 biface manufacturing re-jects (blanks and preforms), 1 fluted point fragment (surface), 1 square-based lance o late point (surface), 3 end scrapers, 6 side scrapers, spurred grav-ers, bend-break tools, knives, scrapers, simple flake cores, disk-shaped bifacial cores, 1 wedge-shaped microblade core, 1 linear flake core, 1 hammerstone, 1 abrader, 10,766 surface flakes, 117,990 excavated flakes

None

None

Rasic 2008, 2011

NR-5 Well-stratified silt and sand layers at the south edge of the hill contained numerous pieces of scattered charcoal. Artifacts were found con-fined to a discrete stratigraphic unit 5–25 cm thick and 30–60 cm below surface.

11 Sluiceway points, 13 uni-facial tools, 30 micro blades, 1 microblade core, 1 sandstone slab, 1 whetstone, 1 hammer-stone fragment, debitage

Anderson 1972; Rasic 2008, 2011

Nat Pass Uppermost lithostratigraphic unit consisted of a gravelly sandy silt that averaged 10 cm in thickness across the excavation block. Nearly all artifacts recovered from the excavation block occurred within this layer, from 0–7 cm below surface. A second stratigraphic layer was discon-tinuous across the excavation block, but where present occurred 7–10 cm below surface. It was a dark brown, organic-rich silt. Several flakes were found within this layer, lying flat and ap-parently in situ. Below this layer was shattered regolith up to 17 cm below surface.

3 Sluiceway projectile point fragments, 26 bifacial blanks and preforms, 1 bifacial knife, 1 oval-platform microblade core, burned debitage, charcoal, burned bone fragments

Rasic 2008Unknown

Irwin Sluiceway Artifacts were shallowly buried but tightly confined to 15–20 cm below the modern surface. Specifics on site stratigraphy have not been presented.

Overlook 15 Sluiceway points (type-specimens), 2 morphologically distinct Sluiceway points (i.e., small, thin, and square-based), 7 bifaces, 3 bifacial knives, 2 biface fragments, 1 unifacial end scraper, minimal debitage

Rasic 2008, 2011; Saleeby 2010

None

Smith et al.

111

Overlook

Overlook

Overlook

Kill/butchery

Nogahabara I Dune Deposits consisted mainly of homogeneous medium-grained sand. Within this occurred a thin discontinuous band of coarse sand and ventifacted granules, differing in color from the rest of the matrix when moist, and potentially representing a paleosol. Artifacts rested upon the potential paleosol where it was visible in the profile.

Notched and thin, concave-based lanceolate bifacial points, bifacial blanks or cores, bifacial preforms, microblades, wide, oval-platform microblade cores and narrow, wedge-shaped microblade cores, burins, core-rejuvenation frag-ments, bifacial knives, unifacial scrapers, blade-like flakes, flake blanks, 5 cobble tools, fire-cracked rock

Bird, mammal Odess and Rasic 2007

Mesa Artifact-bearing sediments were loessy. Below a thin O horizon was a ~15-cm-thick sandy loam that made up the A1 horizon. This was under-lain by a C horizon of gravels and very gravelly loam situated upon degrading bedrock. The soil column was intact with little disturbance from cryogenic processes, with the exception of isolated remnants of ancient frost boils. The Mesa cultural component was associated with the A1 horizon.

154 projectile points/fragments that include the Mesa-type specimens, 212 biface failures, 70 gravers, 13 flake scrapers, 6 unifacial scrapers, 6 informal burins, 237 utilized flakes, 5 hematite pieces (microblade concentration: microblades, micro blade core fragments, 1 wedge-shaped core, related debitage pieces, 1 bead-like item, 3 bifaces, 1 bipolarly fluted biface)

Bever 2000, 2008; Kunz and Reanier 1995; Kunz et al. 2003

Unknown

Unknown

NoneHilltop Artifact distributions suggest that the main site covered approximately 150 m2, and had distinct eastern and western loci. Below a weak modern soil horizon was 20–30 cm of loose sediment overlying shattered bedrock. Artifacts generally occurred within 10 cm of the ground surface.

Mesa points, multi-spurred gravers, bifaces, biface frag-ments, bifacial cores, 1 flake core, flake-blank bifaces, flake scrapers, retouched flakes, burins, 1 unifacial end scraper, 1 large core scraper, biface-manufacture flakes, blade-like flakes, minimal cortical flakes

Bever 2000; Reanier 1995

Putu/Bedwell Geological stratigraphy consisted of an organic mat reaching 8–15 cm in thickness and overly-ing a 23- to 50-cm package of loessy sediment, which in turn rested upon a gravelly regolith. Cultural remains occurred in both the covering soil and loess.

3 fluted-point fragments, 4 non-fluted lanceolate point fragments, several Mesa-type point fragments, 1 triangular point, bifacial blanks, preforms, and failures, as well as numer-ous burins, gravers, scrapers, utilized flakes, cores, blades

Alexander 1987; Bever 2000, 2006b; Reanier 1994, 1995

Engigstciak In general, below a layer of tundra vegetation, a sand to silty sand rich in organics, referred to as a “humic turf,” rested upon (and flanked) marine clays. Cryogenic disturbance to the clays created hummocks and “inter-hummock depres-sions” where interbedded organic layers, clays, and sandy silts occurred. Cryoturbation and solifluction have disturbed much of the site’s archaeological record, but at one locus were intact stratigraphic components. One of these components was the “Buffalo Pit.”

1 bifacial point (fluted on one face) and three unfluted lanceolate points, burins, mi-croblades, end scrapers, side scrapers, pebble pendants, bone needles (not definitely reported from the Buffalo Pit)

Caribou, bison Cinq-Mars et al. 1991; Clark 2009; Hoffecker and Elias 2007; Mackay et al. 1961; MacNeish 1956a,b, 1959, 2000

Geomorphic site setting Depositional setting Lithic assemblage Faunal remains References

Table 6.2 Cont'd.

Biface Traditions of Northern Alaska and Their Role in the Peopling of the Americas

tools and debitage associated with the hearth feature, creat-ing a 13-m2 artifact distribution (Rasic 2008). The lithic as-semblage consists of 24 bifaces, 17 of which are Sluiceway points (Figure 6.2F–G). Flaking debris was uncommon and generally consisted of the by-products of bifacial pressure flaking. The small size of the site and the fact that it con-sisted of a single, hearth-centered artifact scatter suggest that this was a briefly occupied lookout or short-term camp.

Nogahabara INogahabara I is situated in an active dune field in the Noga-habara Dunes of the Koyukuk River valley, ~50 km west of Huslia, northwest interior Alaska. It is reported in Odess and Rasic (2007). Deposits consist of massive sand, and artifacts occurred in a deflated context. The lithic assemblage con-tains complete and unworn or minimally worn formed arti-facts, interpreted as a single-component assemblage repre-

112

0 3cm

A B

E

C

F

D

G

senting a cached or lost toolkit and debris from early-stage core and biface percussion flaking. Raw materials consist of Batza Tena obsidian (originating 140 km to the east), coarse stone, and chert. Small, highly fragmented, and poorly pre-served faunal remains included five fragments of bird bone and one fragment of mammal bone that yielded AMS dates ranging from 13,800–12,565 cal yr BP (Table 6.1). While the horizontal spatial proximity between the dated bone samples and artifact scatter is close, the nature of the association be-tween the dated samples and artifacts is uncertain owing to the lack of stratigraphy (Holmes et al. 2008). A portion of the faunal assemblage was calcined, which, along with the pres-ence of fire-cracked rock, suggested that hearths may once have been present and that the location may have served as a camp.

MesaThe Mesa site (KIR-102) is located ~150 miles above the Arc-tic Circle near Iteriak Creek in the northern foothills of the Endicott Mountains, central Brooks Range. Details on the site and its contents can be found in a number of publications (Bever 2000, 2008; Kunz 1982; Kunz and Reanier 1995; Kunz et al. 2003). The soil column (see Table 6.2) was intact with little disturbance from cryogenic processes and contained a cultural component with 28 hearth features and 154 projec-tile points or point fragments of the Mesa point form (Fig-ure 6.2D–E). The majority of the lithic assemblage consists of both local and non-local cherts, although a few obsidian

flakes derive from the Batza Tena source (Cook 1995). Forty-one AMS dates cluster between 10,300 and 9700 14C yr BP (12,135–11,134 cal yr BP); seven conventional dates were also obtained (Table 6.1). An AMS date of 9330 ± 40 14C yr BP (Beta-125996) was dismissed owing to contamination, while dates of 11,660 ± 80 (Beta-55286) and 11,190 ± 70 14C yr BP (Beta-57430) have been variably interpreted to represent con-tamination, use of old wood, or an earlier occupation (Bever 2000; Hamilton and Goebel 1999; Kunz and Reanier 1995; Kunz et al. 2003; Mann et al. 2001). A microblade concentra-tion associated with a probable, but undated, hearth feature, may represent a younger occupation (Bever 2000, 2008; Kunz et al. 2003). Mesa represents a hunting lookout where hunt-ers prepared and repaired bifacial points.

HilltopThe Hilltop site (PSM-017) is located ~10 km from the con-fluence of the Atigun and Sagavanirktok rivers in the eastern Brooks Range. Initial excavations in the 1970s and a more recent re-investigation are described in Bever (2000) and Re-anier (1995). Excavations determined that artifacts generally occurred within 10 cm of the ground surface. A bulk soil/char-coal sample collected in 1973 yielded a date of 6160 ± 130 14C yr BP (GAK-4924); however, in 1993 Kunz and Reanier col-lected a charcoal sample AMS 14C dated to 10,360 ± 60 14C yr BP (12,513–11,994 cal yr BP) (Beta-69897). The latter date is considered to represent the age of the cultural occupation. Lithic debitage and tools including Mesa-style bifacial points

Figure 6.2 Representative bifacial projectile points from Alaska’s Northern Paleoindian complexes: A–c, fluted-point fragments; D–E, Mesa point fragments; F–G, Sluiceway point fragments. (A–B, Serpentine Hot Springs; c, Putu; D–E, Mesa; F–G, Irwin Sluiceway.)

Smith et al.

113

are predominantly made from local chert sources with no single type dominating. Microblades were initially listed in the 1970 and 1973 catalogs, but Bever (2000) reported these and other debitage pieces missing during his analysis. Tech-nological activities at Hilltop appear to have focused on bi-face manufacture, and the site seems to represent a hunting lookout or short-term camp. As Mesa points represent the main diagnostic artifact in the assemblage, it is assigned to the Mesa complex.

Putu/BedwellThe Putu/Bedwell site (PSM-027) is located in the Sagavanirk-tok River valley, ~45 km north of the continental divide, in the central Brooks Range. The Bedwell locus is near the apex of the knoll while the Putu locus is further downslope on a flat bench (Alexander 1987). Although in the past considered separate sites, here we group them into a single site, follow-ing Bever (2000). Alexander (1987) reported the recovery of four fluted-point fragments on nonlocal materials (e.g. Figure 6.2C), although two are questionable (Bever 2006a), and sev-eral Mesa-type point fragments. The assemblages are domi-nated by local chert, and at one time more than 7000 pieces of debitage existed from the site (Alexander 1987), but most have been lost (Bever 2000). Pieces of bone bear marks from a metal saw attesting to a much more recent age. Combined samples of charcoal yielded six conventional dates that span 15,978–6004 cal yr BP (Table 6.1). Reanier (1994, 1995) AMS dated archived charcoal samples, obtaining ages of 8810 ± 60 (10,158–9631 cal yr BP) (Beta-69901) and 10,490 ± 70 14C yr BP (12,590–12,136 cal yr BP) (Beta-69845), the former thought to be associated with a hearth feature and Mesa point at Putu, and the latter associated with a Mesa point at Bedwell. Bever (2000, 2006a) conducted a spatial analysis of the Putu materials and, finding a random distribution of ma-terials through the site’s deposits, concluded that the buried assemblages represent multiple occupations mixed due to rodent activity, frost churning, and colluviation. Despite the materials coming from a buried context, the coarse excava-tion techniques and early reliance on conventional dating of combined samples of charcoal make it impossible to ascer-tain a valid age range of these cultural occupations, although we interpret them to have occurred before 8000 14C yr BP (8995–8780 cal yr BP), following Reanier (1994) (see also Al-exander 1987; Bever 2000, 2006a; Reanier 1995; West 1996).

EngigstciakEngigstciak (NiVk-1) is situated on a bedrock outcrop located near the mouth of the Firth River, in northern Yukon, 20 km south of the Beaufort Sea (Cinq-Mars et al. 1991; Clark 2009; Mackay et al. 1961; MacNeish 1956, 1959). Stratigraphy at Engigstciak suffered from significant frost heaving and so-lifluction that mixed many of the site’s sediments (Mackay et al. 1961). One locus however, the “Buffalo Pit,” contained relatively undisturbed sediment layers and a lithic assem-blage associated with well-preserved remains of bison, musk ox, and caribou (Table 6.2) (Cinq-Mars et al. 1991). Unfortu-

nately, however, MacNeish did not provide a clear description of the artifacts that derived exclusively from the Buffalo Pit, instead lumping them with other artifacts from other con-texts at the site and calling them the Flint Creek complex. Collagen from a bison tibia, metacarpal, and metatarsal, all with cut marks, yielded dates with fairly large standard errors that span 11,989–9961 cal yr BP (Table 6.1) (Cinq-Mars et al. 1991). These were associated with a large number of lanceo-late-shaped projectile points that MacNeish (1959) described to resemble Milnesand, Plainview, and Angostura forms. The directly dated faunal remains from Engigstciak’s Buffalo Pit provide clear evidence that humans hunted and processed large game in Arctic Beringia during the earliest Holocene.

Additional SitesTwo other sites deserve brief mention here, despite their lack of 14C dating or final reporting: Batza Tena and Girls Hill. The Batza Tena obsidian source is located at the head of the Koyu-kuk lowlands where more than 50 archaeological sites, 10 of which produced 18 fluted points, have been discovered since 1971 (Clark and Clark 1993). Chronological investigations failed to produce meaningful results (Clark and Clark 1993; Hamilton and Goebel 1999); however, the evidence of fluted-point morphology and technology present in the Batza Tena assemblages is important and considered in the sections that follow. Likewise, Girls Hill, located along the Jim River in the southern foothills of the Brooks Range, yielded four fluted points from a context with no reliable chronological control (Dumond 1980; Gal 1976), but lack of reporting makes it dif-ficult for us to interpret these finds.

DiscussionAs is evident from the above presentation, 11 dated archaeo-logical sites help document the terminal Pleistocene/earliest Holocene record of humans in northeastern Beringia. Archae-ologists studying these sites conventionally group them into three complexes based on diagnostic bifacial points—the fluted-point, Sluiceway, and Mesa complexes. In the discus-sion below we review what is known and currently inferred about these three complexes, focusing on chronology, bifa-cial-point technology, lithic technology, subsistence, and set-tlement organization. We also attempt to put the Northern record into the context of Beringian and early American pre-history, by considering its relation to Paleoarctic microblade industries in Beringia and Paleoindian bifacial industries in temperate North America.

GeochronologyAt least 11 archaeological sites provide geochronological data regarding the terminal-Pleistocene/earliest-Holocene cultural record of northern Alaska and neighboring Yukon. In Figure 6.3 we present the summed probabilities of calibrated dates for each of these sites, based on the calibrated dates in Table 6.1 (OxCal 4.2) and resulting summed probabilities using Ox-Cal’s Sum function. The result is a graphic representation of the likelihood of ages for the various cultural occupations.

Biface Traditions of Northern Alaska and Their Role in the Peopling of the Americas

114

Four sites—Nogahabara I, Mesa (locality B), Putu, and Tuluaq Hill—may contain pre–Younger Dryas (i.e., pre-12,900 cal yr BP) occupations; however, only Tuluaq Hill is firm. Nogahabara I’s ~13,600 cal yr BP date is coupled with a much younger age of ~12,700 cal yr BP, and both come from a deflated dune setting (Odess and Rasic 2007). Holmes et al. (2008) understandably question the association of these dates with artifacts, and we agree that the dating of Nogahabara 1 is inconclusive. Mesa has two charcoal dates of ~13,500 and ~13,100 cal yr BP (Kunz and Reanier 1994, 1995; Kunz et al. 2003), and although site excavators do not dismiss them as potentially representing the earliest human occupation of the site, others have questioned their accuracy given the half-mil-lennium age span for this single hearth feature and the site’s 41 other dates that cluster between ~12,100 and ~11,100 cal yr BP. An alternative interpretation is that the old dates represent the burning of old wood in a much younger hearth

Hill seem secure. Six AMS 14C dates on hearth charcoal, all from the same stratum, indicate an age for this occupation of 13,250–12,900 cal yr BP (Rasic 2008, 2011). This, then, is the strongest candidate for a late-Allerød occupation of northern Alaska. Most early northern Alaskan occupations date to the Younger Dryas and the five centuries following it, 12,900–11,200 cal yr BP. Among them are the fluted-point assem-blages at Raven Bluff and possibly Serpentine Hot Springs, the Mesa lanceolate-point assemblages at Hilltop, the four Mesa site loci and possibly Engigstciak, and the Sluiceway lanceolate-point assemblage at Irwin Sluiceway and likely NR-5 and Nat Pass (Figure 6.3). Mann et al. (2001) pointed out that some individual hearth features at Mesa have spans ap-proaching 1300 calendar years; they argue that this, coupled with the near 1500-year span for the entire site occupation, could be the result of the Younger Dryas radiocarbon plateau.

Figure 6.3 Summed probability distributions of radiocarbon dates from the Northern Paleoindian occupations described in text (calibration and summed probability analyses conducted using OxCal 4.2.2 Bronk Ramsey [2013]; r:5 Atmospheric data from Reimer et al. [2009]). Note that NR-5 dates are geological dates only and roughly provide limiting ages for the Sluiceway component.

(Hamilton and Goebel 1999; Mann et al. 2001). The 11,470 14C yr BP date from Putu has a substantial 2600 calendar-year standard deviation, a questionable relationship to cultural activity at the site, and could not be replicated with AMS dat-ing of archived charcoal (Reanier 1995). Reanier (1994, 1995) suggested the material dated may in fact be a pocket of nat-urally buried vegetation that only superficially resembled a hearth. Unlike these other sites, the early ages from Tuluaq

This plateau, though, persisted from 10,600 to 10,300 14C yr BP (or 12,700–11,900 cal yr BP) (Fiedel 2011), while most Mesa occupations seem to post-date 11,900 cal yr BP. A few dates suggest occupations extending past 11,200 cal yr BP. Serpentine’s fluted points could date to as late as 10,700 cal yr BP, but as Goebel et al. (2013) argue, the young-est dates are on charcoal of low-lying shrubs, not willow, and could be unrelated to human occupation of the site. The

Smith et al.

115

young date from the East Ridge locality at Mesa was consid-ered by site excavators to be a contaminated sample (Kunz and Reanier 1995), and deleting it from our analysis tight-ens the Mesa distributions considerably. Engigstciak’s broad calendar-year range is the product of three dates with large standard errors. These dates are from three bones with clear cut marks related to cultural activities that overlap at 1 sigma. When they are averaged, a weighted-mean of 9770 ± 111 14C yr BP is obtained, a somewhat tighter 11,592–10,721 cal yr BP. The late Putu dates are as problematic as the early Putu dates, given the problems with the site’s context and excava-tion protocols discussed by Bever (2006b) and Reanier (1995). The fluted points at Girls Hill must be considered undated as the pre-AMS dates derive from combined charcoal samples that were not collected from specific, identifiable cultural features, and derive from a shallow stratigraphic context con-taining abundant natural charcoal typical of the site’s boreal forest setting (Gal 1976; R. Gal, pers. comm.). The chronological evidence at hand suggests that the Paleo indian bifacial complexes of northeastern Berin-gia disappeared by 10,700 cal yr BP, perhaps earlier. More specifically, the Sluiceway complex spans ~13,200–11,000 cal yr BP, the fluted-point complex spans ~12,700–10,700 cal yr BP, and the Mesa complex spans ~12,500–10,700, if the weighted mean for Engigstciak is included. Therefore, whether they ultimately differed functionally or culturally, all three complexes potentially existed simultaneously and in-explicably disappeared within a few hundred years of each other. Paleoecological records, however, indicate significant warming and an increase in precipitation after 11,000 cal yr BP, suggesting the development of more modern tundra con-ditions that may relate to change in human adaptation (Mann et al. 2001).

Bifacial Point Technologies The geographically widespread presence of three diagnostic bifacial-point forms makes northern Alaska’s early bifacial technology potentially distinct from coeval technologies common to the interior river valleys (Dixon 1985; Dumond 2001; Goebel et al. 1991; Potter 2011; Powers and Hoffecker 1989; but see Hoffecker 2001; Holmes 2001). Often com-pared with Paleoindian tool forms of mid-continent North America, northern Alaska’s bifacial projectile points, as well as their condition at time of discard, demonstrate a focus on formal tools, procurement and transport of high-quality raw materials, and strong reliance on projectile weaponry (Bever 2000; Clark 1991; Kunz et al. 2003). Despite these robust commonalities, each point form is morphologically and tech-nologically distinct (Figure 6.2). Their diagnostic attributes, however, fail to serve as precise chronological markers be-cause, as demonstrated above, their timespans overlap con-siderably. Geographically, fluted points occur in sites across northern Alaska and Yukon Territory. Sluiceway points appear to be more narrowly distributed and concentrated in the western Brooks Range, while Mesa points are concentrated in the central Brooks Range. The distributions of Mesa and

Sluiceway points overlap considerably, and they occasion-ally co-occur. For example, 14 “Type A” bifaces are noted at the Mesa site, and these fall squarely in the range of shape and size variation for Sluiceway projectile points (Kunz et al. 2003:29; Rasic 2008). Fluted-point forms recovered from Serpentine Hot Springs, Raven Bluff, Girls Hill, Putu, and Batza Tena have relatively thin cross sections and widths averaging 2–3 cm. Preforms from Raven Bluff and Batza Tena suggest that semi-regular pressure flaking was often imposed on a flake blank to create a center ridge down the long axis of the finished point. Basal cavities are characteristically deep and fashioned into a distinctive V-shape (Clark 1991; Clark and Clark 1993; Goebel et al. 2013). Basal cavity depth was increased through the removal of multiple fluting flakes that thinned the entire proximal face, usually removing evidence of invasive lateral flaking. Points were finished with fine pressure flaking and ground along the lateral and basal margins, creating straight lateral edges and a refined basal cavity. Sluiceway projectile points from Irwin Sluiceway, Tuluaq Hill, NR-5, and Nat Pass are distinctively large in size and nearly always have rounded, convex bases that were not ba-sally thinned. The manufacturing process began with percus-sion-flaked bifacial blanks (Rasic 2008). Finished points were shaped with robust, semi-regular, collateral pressure flaking that produced thick biconvex cross-sections preserved along their entire long axes. Plan outlines are oblanceolate, with ex-curvate lateral edges and heavily polished, or ground, proxi-mal margins (Rasic 2011). Mesa projectile points found at Mesa, Hilltop, Engigst-ciak, Putu, and Bedwell are collaterally pressure flaked to cre-ate a distinctive medial ridge down the long axis of the point (Kunz et al. 2003). The reduction sequence reconstructed from the Mesa site assemblage demonstrates that these points began as bifacial blanks manufactured from tabular nodules of chert (Bever 2000). Remnant percussion scars from early stages of production are often still present on finished points (Rasic 2011). Mesa points are diamond to lenticular in cross section, fairly thick (average 7–10 mm), and basally thinned, forming a slightly concave base with distinct cor-ners (although slightly convex-based forms also occur) (Bever 2000; Kunz 1982; Kunz and Reanier 1995). Lateral and basal margins are heavily edge-ground (Bever 2000). Despite the differences in form and basal treatment, the three projectile point forms display several common features such as the creation of a midline ridge down their long-axis and use of high-quality toolstone in their production. While this is not unique to assemblages in interior northern Alaska where high-quality chert is abundant, the Batza Tena obsid-ian at Serpentine Hot Springs, NR-5, Putu, and Mesa dem-onstrates that raw material was often transported hundreds of kilometers. An almost characteristic consistency in edge grinding and refinement, their fragmentary nature, and pres-ence of impact fractures confirms their principal use as hafted projectile points (Goebel and Potter, in press). Technological similarities between Sluiceway and Mesa

Biface Traditions of Northern Alaska and Their Role in the Peopling of the Americas

116

projectile points in particular have been noted (Bever 2001, 2006b; Mann et al. 2001; Rasic 2011). Both are lanceolate shape, with thick cross sections, heavily ground proximal margins, and robust, sub-regular collateral pressure flaking. They were also often rejuvenated in their hafts after break-age for continued use as projectiles. There are, however, important distinctions between these two kinds of projec-tile points. Only Sluiceway projectile points appear to have served as multifunctional tools, at times being recycled into hafted knives with sharp-edged but rounded distal ends and use-wear polish indicative of cutting (Rasic 2008:270). Mesa and Sluiceway points also derived from distinct produc-tion trajectories. Sluiceway points resulted from a lengthy, multistage shaping process that included extensive pres-sure flaking of large, percussion-flaked bifacial blanks (Ra-sic 2008:249), while Mesa projectile points were produced from a shorter reduction sequence that began with a smaller blank and involved much less extensive pressure flaking. The less extensively shaped nature of the smaller Mesa projectile points demonstrates that they are not simply resharpened or recycled versions of Sluiceway projectile points. There are sites at which the different point technologies co-occur, prompting questions about the precise nature of their relationship. At most sites, unfortunately, the chrono-logical and stratigraphic control is not sufficient to address this issue. At Putu, where there are good examples of both fluted and Mesa projectile points, the palimpsest nature of the shallow deposit leaves the exact relationship between these artifacts unresolved. A single fluted point was discov-ered at Tuluaq Hill that is otherwise dominated by Sluiceway technology, but it was a surface find that cannot be tied to dated components of the assemblage. Misidentification of artifact type is also an issue. For example, we question the validity of the “flute” MacNeish (1956, 1959:96,55, pl. I #5) described on a point from Engigstciak, which is also from uncertain context outside the Buffalo Pit. Likewise, at the Mesa site, a few Mesa points have also been described as fluted (Kunz et al. 2003), but we are inclined to classify these as basally thinned rather than purposefully fluted. One pos-sibly strong association among two of the projectile point types does come from the Mesa site, however. As mentioned above, in addition to 154 Mesa points, investigators also report 14 “Type A” bifaces that we would classify as Sluice-way projectile points (Kunz et al. 2003:29). It is important to note that both types occur within the same well-dated, hearth-centered artifact clusters and a meaningful, and rather precise, association is indicated. It seems reasonable to con-sider that Mesa and Sluiceway points may represent func-tionally distinct tools found within one broad technological repertoire. The specific nature of the functional distinction is unknown but could relate to different prey species, hunt-ing tactics, or projectile propulsion methods (e.g., dart vs. hand-held thrusting spear). Whether fluted points form yet another component of this technological repertoire is much less clear given the ambiguous contexts in which most fluted projectile points have so far been found.

Lithic Technological OrganizationThe early archaeological sites reviewed here contained ad-ditional tools and debitage potentially associated with the projectile points; however, contextual shortcomings of most of these sites often make it difficult to prove they represent the same occupations rather than palimpsests. Nonetheless, we endeavor to make some inferences about the overall char-acter of the lithic technologies represented. It is no surprise that high-quality raw materials, primarily chert, dominate all of these assemblages from northern Alaska because the re-gion, particularly the western Brooks Range, contains exten-sive limestone and chert formations and toolstone is avail-able in many primary and secondary sources (Mull 1995). For the fluted-point sites, high-quality raw materials, pri-marily cryptocrystalline silicates, dominate assemblages (Al-exander 1987; Bever 2006a; Clark and Clark 1993; Goebel et al. 2013; Hedman 2010). Raven Bluff, Putu, and the Batza Tena sites are located near sources of high-quality toolstone and have assemblages with high percentages of the local ma-terial. While local raw materials dominate the Putu assem-blage, the two fluted points are made from materials uncom-mon to the assemblage, one of which is Batza Tena obsidian (> 400 km southwest) (Alexander 1987). This is in contrast to the Serpentine assemblage, which is dominated by non-local raw materials (95%), including obsidian from Batza Tena (450 km east) and chert and chalcedony not local to the area but likely accessible in rivers emerging from the western Brooks Range (250 km north) (Malyk-Selivanova et al. 1998). Simple flake cores and bifacial cores were found at Batza Tena and Putu, and blade/bladelet cores were recovered from some of the fluted sites, as were blades or bladelets (Bever 2006a; Clark and Clark 1993; Goebel et al. 2013). Bifacial blanks and preforms provide evidence of formal tool production at Raven Bluff, Putu, and Batza Tena, while the assemblage at Serpentine contains only finished or nearly finished bifaces (Alexander 1987; Bever 2006a; Clark and Clark 1993; Goebel et al. 2013; Hedman 2010). The same patterns are seen in the debitage: most fluted-point assemblages have a combination of core-reduction and biface-reduction debitage, while Serpentine has only fine secondary debitage, including a significant number of channel flakes. Scrapers are common, being invasively re-touched, and again, often made on non-local materials (but not at Batza Tena) (Clark and Clark 1993; Goebel et al. 2013). Flake tools in contrast are relatively rare; however, spurred gravers made on biface-thinning flakes of non-local stone were present at Serpentine and Putu. Burins have been reported from Putu and surface contexts at Batza Tena localities (Alexander 1987; Clark and Clark 1993). At Putu, however, these artifacts were later shown to be snapped flakes that only superficially resem-bled burins (Bever 2006b), and no burins have been found in the sealed stratigraphic contexts at Serpentine or Raven Bluff (Goebel et al. 2013; Hedman 2010), or in regular association with the many surface finds of fluted points across northern Alaska. Burin technology does not seem to be associated with fluted-point technology. The formalized nature of the assem-blages, focus on hafted bifaces, and inter-site transport of tools

Smith et al.

117

together suggest a high degree of mobility, and the varied site types represented (i.e., tool-production sites and short-term camps or lookouts) more specifically suggest logistically orga-nized mobility. High-quality cryptocrystalline silicates dominate the Mesa complex sites of Mesa, Hilltop, Engigstciak, Putu and Bedwell (Bever 2000; Kunz et al. 2003; MacNeish 1956; Re-anier 1995). All are near sources of high-quality toolstone, and the assemblages include primary production debris such as bifacial blanks, preforms, and bifacial-production debris and multi-spurred gravers providing evidence of the produc-tion of formal tools. Bifacial and flake cores were found at Hilltop, while informal cores of local tabular chert were pres-ent at Mesa and Putu (Alexander 1987; Bever 2000, 2006a; Kunz et al. 2003). The use of local, high-quality raw materials in the manufacture of hafted bifaces, gravers, and scrapers suggests formal tool manufacture took place at Mesa sites (Bever 2000; Kunz et al. 2003; MacNeish 1956). Sites posi-tioned near sources of good raw material, yet where formal tools were used, suggest that while Mesa complex peoples were likely highly mobile, their logistical movements may have varied according to season or predicted animal move-ments. High-quality raw materials also dominate the Sluiceway assemblages (e.g., Malyk-Selivanova et al. 1998). Tuluaq Hill and Nat Pass are located immediately adjacent to the sources of toolstone present in their assemblages, although some non-local material use is also indicated, e.g., Batza Tena obsidian found at NR-5, and Tuluaq Hill. These sites also have evidence of biface production (e.g., bifacial blanks and preforms in vary-ing stages of manufacture) (Rasic 2008). Bend-break tools, hy-pothesized to facilitate preparation of wood or antler handles and foreshafts, were found at Tuluaq Hill and Nat Pass (Rasic 2011). The Sluiceway sites discussed here all possess evidence of a formal bifacial technology focused on the manufacture and maintenance of Sluiceway projectile points, as few other formal tools occur at every site. Scrapers present at each site, except Nat Pass, suggest animal processing occurred as well, but these tools may have served in woodworking, as the bend-break tools suggest (Rasic 2008). This evidence and the strikingly large number of Sluiceway projectile points and frag-ments at each site support Rasic’s (2008) hypothesis that these sites often represent repeated use of key locations over many seasons and were likely created by small hunting groups gear-ing up for intercept hunting of caribou. While more variability is present in the fluted and Mesa site types, all three complexes have sites ideally situated near high-quality toolstone and provide broad views of the sur-rounding landscape. Mobile prey and seasonal climatic ex-tremes represent risk factors that ultimately led to formal tool kits and logistical mobility organized within potentially long-distance seasonal rounds.

Microblades and Northern PaleoindiansDecades of research have yet to confirm a confident associa-tion between microblades and Northern Paleoindian bifacial

technology (Bever 2008; Clark 1984; Rasic 2011). Evidence of microblade technology, however, is present to varying de-grees at many of the sites. The question remains, though, whether the presence of microblades represents a mode of reduction complementary to formal biface manufacture or separate assemblages left behind by technologically unre-lated peoples (Dixon 2001; Hoffecker 2005; Holmes 2001; Potter 2008). If microblade technology does form part of the arctic Pa-leoindian technical repertoire, an important consideration is the kind of microblade core technology utilized—is it spe-cially designed wedge-shaped-core-and-microblade technol-ogy (common in central Alaska), or is it part of a more casual bladelet technology grading into specimens metrically typed as microblades? The Serpentine fluted points occurred with a few bladelets and microblades, and a platform tablet from a conical-shaped core was found isolated in the western area of the excavation where its association with the Paleoindian component is questionable, but so far no clear wedge-shaped microblade cores have been found (Goebel et al. 2013). Cur-rently, we cannot conclude that formalized wedge-shaped-core-and-microblade technology was utilized there. Wedge-shaped cores were reported from Girls Hill and Batza Tena (Clark and Clark 1993; Gal 1976), but their association with fluted points and datable material is tenuous. Bladelet and microblade cores were reported from Putu (Alexander 1987); however, no formally prepared microblade/blade cores exist in the collection and widths of detached pieces form an ir-regular distribution with no evidence of standardized pro-duction. Microblades, a wedge-shaped microblade core, and microblade-core technical spalls were also recovered at the Mesa site, but Bever (2008) and Kunz et al. (2003) consid-ered them to be unrelated to debris more characteristic of the Mesa bifacial complex, based on their distribution and in-corporation of a maroon-gray chert not present in any of the Mesa complex artifacts. The presence of a Mesa-type biface that was recycled into a microblade core also suggested to Kunz et al. (2003) that they resulted from a later occupation. Interestingly, a maroon-gray chert biface, fluted from both ends on both faces, was assigned to the microblade compo-nent (Bever 2008; Kunz and Reanier 1995) and is thought to represent an expedient exercise in microblade production on a biface (M. Kunz, pers. comm.). We hasten to add, however, that the microblade concentrations at Mesa remain undated and there is a real possibility that the microblades are part of the Mesa occupation. Concerning the Sluiceway complex, clear evidence of micro blades and microblade cores comes from Tuluaq Hill, NR-5, and Nat Pass, but the latter two sites yielded oval-plat-formed cores, while the only wedge-shaped core was found on the surface at Tuluaq Hill (Anderson 1972; Gal 1982; Rasic 2008). The stratigraphic context of these sites precludes de-finitively associating (or dis-associating) the micro blade tech-nology from the bifacial component of these assemblages. While not clearly dated, the assemblage from Nogahabara I

Biface Traditions of Northern Alaska and Their Role in the Peopling of the Americas

118

contains lanceolate projectile points and both oval-platform and wedge-shaped cores. If this assemblage does in fact rep-resent a single cultural component as Odess and Rasic (2007) suggest, a clear association between these technologies is evident there. Differences in the method of microblade manufacture represented at the sites is an aspect of variability that may be useful in addressing two major issues concerning the role micro blades played in northern Paleoindian adaptations. First is whether true microblades and wedge-shaped core manufacture can be unequivocally associated with northern Paleoindian bifacial technology. Thus far, evidence consists of palimpsest situations (Batza Tena), surface contexts (Tuluaq Hill) or segregated clusters of artifacts (Mesa) (Bever 2008; Clark 1991; Clark and Clark 1993; Gal 1976; Kunz et al. 2003; Rasic 2011). Second, conical or oval-platformed microblade cores found at Nat Pass, and a conical/oval-shaped core tab-let at Serpentine, potentially produced a graded bladelet-to-microblade continuum, significantly different from the stan-dardized micro blades produced through wedge-shaped cores at terminal-Pleistocene/early-Holocene Denali complex sites in central Alaska and Diuktai complex sites in northeast Asia. When considering Sluiceway sites, in which both bifacial and flake cores are found, Rasic (2011) recognizes a poten-tial combination of bifacial and microblade technology that may have resulted from differing prey species, seasonality, or raw-material constraints (Morlan 2003; Rasic 2008). Certainly in these and other Northern Paleoindian assemblages micro-blades are present, and if they actually represent a technolog-ical association with the strong bifacial industries of north-ern arctic Beringia, they may ultimately speak to an increased technological flexibility of these Northern people, and fur-ther, possible historical interaction of Paleoindian groups and cultural transmission of technologies that originated in both the New (Plains Paleoindian bifacial and fluted technolo-gies) and Old (Northeast Asian microblade complexes) Worlds (Bever 2001; Hoffecker et al. 1993; Kunz and Reanier 1995; Sellet 2001; Slobodin 2001, 2011).

Subsistence OrganizationFaunal remains are rare in northeastern Beringia’s archaeo-logical sites, primarily because of shallow deposits and poor preservation potential. This equates to a poor record of subsistence organization. It is no coincidence that the two most deeply buried sites in the study area—Raven Bluff and Engigstciak—also have produced the best-preserved faunal assemblages. For the fluted-point sites, signs suggest that subsistence in part focused on caribou (Rangifer tarandus). The strongest ev-idence of this comes from Raven Bluff, where first reports char-acterize the fauna as chiefly caribou (Hedman 2010). Likewise, identifiable specimens at Serpentine Hot Springs are from caribou/deer–sized mammals, despite a highly fragmented and carbonized assemblage (Goebel et al. 2013). Unspecified birds and/or small mammals are also present in these two as-semblages—both at Raven Bluff and a hare-sized mammal at

Serpentine. With the evidence on hand, we cannot judge the relative importance of these animals in early human diets. Sites of the Sluiceway complex unfortunately have not yielded identifiable faunal remains; however, based on high numbers of Sluiceway points in sites adjacent to intercept points along caribou migration routes, Rasic (2011) infers that early humans in the uplands of the western Brooks Range and Noatak basin targeted aggregated caribou during fall and/or spring migrations. During the rest of the year, he argues, these same humans may have practiced a more gen-eralized subsistence pattern, taking bison, sheep, musk oxen, and dispersed caribou. No identifiable faunal remains were recovered from the Mesa complex’s type-site, nor were they found at Hilltop and Bedwell. Only Engigstciak, a possible Mesa complex oc-cupation, produced well-preserved animal bones, primarily steppe bison (Bison priscus) as well as caribou and musk ox. It is difficult to estimate the relative importance of bison, or any other single species, in the subsistence economies of early inhabitants of the region owing to such a small sample size. Paleontological finds from the North Slope show that bison, as well as horse, persisted until at least 12,500 cal yr BP (Mann et al. 2013) and clearly overlapped with human oc-cupation of the region. These now-extinct species co-existed with extant large mammals such as caribou, moose, musk ox, and sheep, and presented an especially rich range of human subsistence opportunities. This diverse community of large game was a fleeting and unique event in the paleoecological history of the region with key representatives of the earlier Pleistocene fauna and later-Holocene fauna coexisting. This ecological backdrop may well have structured land-use pat-terns in ways fundamentally different from those seen in re-cent hunter-gather groups in the Far North for which caribou has served as a central economic focus.

Settlement OrganizationWe base our discussion of settlement organization mainly on site-location information and assemblage structure. Most of the sites discussed here are in lookout situations on knobs or ridges 15–60 m above surrounding valley bottoms and in locations that provide good views of the surrounding landscape. Artifact assemblages typically contain a narrow range of items dominated by hunting weaponry, and the sites lack evidence for housing or storage features. As such, they have been interpreted as very briefly occupied, spe-cial-purpose locations functioning as hunting lookouts or short-term hunting camps. A few sites such as Batza Tena, Raven Bluff, Putu, Bedwell, and Tuluaq Hill are located adja-cent to lithic raw-material outcrops or plentiful secondary sources and also functioned as lithic workshops. The only sites that do not fit the overlook pattern are Nogahabara I and the Buffalo Pit at Engigstciak, the former seemingly rep-resenting a short-term camp in a dunefield (Odess and Rasic 2007), and the latter possibly representing a kill/butchery site (Cinq-Mars et al. 1991). The variability in site types suggests a pattern of logistical mobility for Northern Paleo-

Smith et al.

119

indians; however, only a narrow and certainly incomplete range of activities is represented in the current sample of sites. The most glaring gap is residential sites. Perhaps base camps did not exist, but more likely their apparent absence reflects the vagaries of site discovery and preservation. It is likely that residential camps were closely associated with some of the repeatedly used lookout locations, and prob-ably located nearby on valley bottoms close to water, and shrub thickets that provided woody fuel, building materials, and shelter from wind. Rasic (2011) presents an interesting analysis of assem-blage structure, finding that Sluiceway sites represent either generalized short-term hunting camps (with multiple tool forms) or task-specific intercept lookouts (with large propor-tions of hafted bifaces) in upland settings seasonally visited by large caribou herds. Otherwise, studies of assemblage structure focusing on site-occupation spans have not been conducted. Suffice it to say that a high-mobility settlement strategy, as discussed above, can be inferred from the re-peated presence of

1) biface-focused industries,

2) use of bifaces as cores,

3) use of bifacial thinning flakes as tools, and

4) a focus on high-quality lithic resources and their inter-site transport.

Analogues and ConnectionsThe origins of the Northern Paleoindian bifacial complexes are not well understood; however, most would agree that they are rooted in North America, not northeast Asia. Fluted points have never been found in the Russian Far East; the “flute” on the biface from Uptar-1 (King and Slobodin 1996) is likely the result of severe impact damage, not intentional basal thinning (Goebel et al. 2013; Meltzer 2009:189). Simi-larly, bifacial points similar to Sluiceway and Mesa forms have not been found across the Bering Strait, and no sites in neigh-boring Chukotka have been unequivocally dated to the late Pleistocene or earliest Holocene (Goebel and Slobodin 1999; Slobodin 2011). Connections with Beringian industries in central Alaska are not well founded, either. Hoffecker (2011; see also Hof-fecker and Elias 2007) has suggested that Mesa points oc-cur in some Denali complex industries, and Holmes (2001) has subsumed the northern Alaskan sites into a seemingly pan-Alaskan “Beringian tradition.” While there are indeed ex-amples of lanceolate-shaped projectile points in some early assemblages from interior Alaska, such as the Denali com-plex’s convex-based points that lack basal thinning, we see little in common in the bifacial technologies to agree with these assessments. The distinctive features of the northern bifacial projectile point technologies are technological rather than morphological, and include lengthy bifacial production sequences, extensive pressure flaking, and consistent edge grinding. Also of note are assemblage-scale patterns that contrast strongly between interior and northern Alaska. Bifa-

cial projectile points are a major component of most North-ern assemblages, with some sites containing well over 100 projectile points suggesting high levels of standardization. In contrast, interior assemblages rarely have more than a dozen projectile points, and among these small samples there is considerable formal and technological variation (Cook 1969; Hamilton and Goebel 1999; Holmes 2001; Powers and Hof-fecker 1989). The underlying explanation for these regional differences is unresolved. It may be that fundamental ecologi-cal differences between northern and interior Alaska engen-dered different subsistence and economic adaptations and different ways of organizing technological systems. We argue that the bifacial industries highlighted here are unique for Beringia and may represent a distinct Beringian tradition of tool production (following Kunz et al. 2003). Are the origins of the Northern Paleoindian industries rooted in northern Alaska (Kunz et al. 2003), or in temper-ate North American late-Paleoindian complexes (Bryan 1969; Dixon 1999; Haynes 1987)? The answer to this question is not simple. Sluiceway points at Tuluaq Hill are too old to have originated from a Plains context, and if they are a compo-nent of the Mesa complex at the Mesa site as proposed here, then this implies a local origin for both Sluiceway and Mesa points. However, the fluted points of Serpentine and Raven Bluff are consistent, technologically and chronologically, with post-Clovis fluted technologies of temperate North America (e.g., Folsom), and together with similar fluted points from northeast British Columbia suggest the presence of a north-ern Plains Paleoindian tradition that spread northward into Alaska at the end of the Pleistocene (Bever 2001; Dumond 2001; Ellis et al. 2011; Goebel et al. 2013; Hoffecker and Elias 2007; Willey 1966; Wormington 1957). One of us (Smith) is currently conducting an interregional technological and mor-phological analysis of late-Paleoindian fluted points to in-vestigate possible connections with various late-Paleoindian technologies of the North American Plains, Mackenzie Corri-dor, Canadian Maritime Provinces, Great Lakes area, and New England.

conclusionsBased on the above review, we offer the following conclu-sions about the Northern Paleoindian complexes of north-eastern Beringia:

1) Given current evidence, Northern Paleoindian occupa-tions span a 2500-year period, from ~13,250 to 10,700 cal yr BP, from the late Allerød through Younger Dryas and the 500–1000 years following. With continued geochronological research at sites like Serpentine and Engigstciak, this span of time could become tightened considerably, especially on the young end, by as much as 500 years.

2) Northern Paleoindian industries can be organized into three complexes based on distinctive forms of bifacial points: fluted, Sluiceway, and Mesa. These complexes, however, are best considered as heuristic tools for shar-

Biface Traditions of Northern Alaska and Their Role in the Peopling of the Americas

120

ing information and posing new questions about the early archaeological record of this region. Projectile points are an especially recognizable aspect of these assemblages and were perhaps an important part of prehistoric technological systems, but they are only one facet of a broader and undoubtedly complex and sophisticated technological repertoire that included a wide variety of stone and organic tools. Given that these distinct kinds of projectile points were synchro-nous for much of their durations and that in several cases they co-occur within single sites, it may be inac-curate to assume that they represent different cultural traditions or human groups.

3) Fluted, Sluiceway, and Mesa points are lanceolate in shape and generally produced through “to-the-mid-line” flaking, morphologically creating thick biconvex to diamond-shaped cross sections. Fluted projectile points, however, were derived from less extensively and less regularly flaked, thinner preforms. Treatment of bases differed, too, with fluted points having deeply concave bases produced by removing multiple channel flakes, Sluiceway points having convex bases, and Mesa points typically having shallowly concave bases.

4) The Northern Paleoindian lithic industries were based primarily on high-quality toolstones, for example, locally available cryptocrystalline silicates, and occa-sionally non-local obsidian, chiefly from Batza Tena. Bifacial as well as unifacial tools, such as small gravers often made on biface-thinning flakes, were made from raw materials transported hundreds of kilometers from source locations.

5) A firm understanding of the relationship between Northern Paleoindian bifacial industries and micro-blade industries still eludes us. Where specially de-signed wedge-shaped cores have been found, they are from shallow contexts that may represent palimpsests, although in a few cases the two technologies have been found together in isolated and close spatial proximity. These surface contexts, however, are undated. In other cases, where microblades but not wedge-shaped cores are present, microblades may represent a continuous reduction system with bladelets and possibly blades, and appear to have been detached from less well-pre-pared conical/oval or columnar cores.

6) Only three sites provide direct evidence of subsistence activities, and they indicate large-ungulate hunting, pri-marily of caribou but also bison and musk oxen. How much early hunters relied on bison or other large mam-mals present on the landscape at this time remains an open question.

7) Most Northern Paleoindian sites are in lookout situa-tions, and so far no obvious residential site has been found. The combination of the transport of tools on high-quality raw materials (sometimes hundreds of kilometers) and the creation of formal tool assemblag-

es together suggest a highly mobile settlement strat-egy that may have been seasonally driven.

Explanations for the lithic variability present among the Northern Paleoindian complexes remain beyond our reach. Possible explanations include: variation in hafting method (e.g., Dixon 1999; Eerkens and Lipo 2005); species predict-ability and/or size of prey (Churchill 1993; Kelly and Todd 1988; Torrence 2002); seasonality (Bamforth 1991; Binford 1980), or in the end, culture history (Dixon 2011; Goebel et al. 1991; Hoffecker 2001, 2011; Kunz and Reanier 1994). De-tailed comparative analyses are needed to integrate techno-logical and geometric morphometric approaches. Likewise, future research needs to address the proposed connection between the Northern Paleoindian complexes and those from temperate North America. On the one hand, if the dates for the early Sluiceway complex occupation at Tuluaq Hill are viable, as we argue, and if Sluiceway points also occur at Mesa 1000 years later, this implies that a Plains origin for both the Sluiceway and Mesa complexes is un-likely. On the other hand, new geochronological evidence from Serpentine and Raven Bluff suggests that the North-ern fluted-point complex is too young to represent a Clo-vis ancestor (Goebel et al. 2013; Hedman 2010), while their ages, bifacial-point forms and technologies suggest closer links with late-Paleoindian fluted complexes in southern Canada or the northern Plains (Goebel et al. 2013). These are links too specific and synchronous to have resulted from independent invention. These similarities strongly suggest a northward spread of late Paleoindians, or at least late-Paleoindian fluting technologies, into the interior Ice-Free Corridor during the Younger Dryas. Recent genetic studies have found that Plains bison dispersed northward into the widening Ice-Free Corridor ~12,500 cal yr BP (Shapiro et al. 2004; Wilson et al. 2008), and perhaps fluted-point makers followed them to the south-eastern edge of Beringia. From there either they continued to disperse (or their technology continued to diffuse) into northern Yukon and Alaska, eventually shifting subsistence toward caribou, either by simply expanding beyond the range of northern bison (Rasic and Matheus 2007) or by meeting a new adaptive challenge as grasslands gave way to ecologi-cal settings in which caribou thrived (Mann et al. 2001). Ulti-mately, arctic Paleoindians were adaptively equipped to reach Serpentine Hot Springs and Raven Bluff in far northwestern Alaska by 12,000 cal yr BP. The peopling of Beringia was obviously a much more complicated process than the traditional, unidirectional Ber-ing land bridge theory for Native American origins allows, and variability in arctic Beringia’s bifacial industries is further compounded by unequivocal evidence of fluting technology in Alaska at the end of the Pleistocene. This variability, how-ever, may provide new means to investigate the mechanism for transmission of Paleoindian technology across the north and assess relationships between early northern Beringian industries and coeval traditions utilized by Paleoindians throughout Alaska and temperate North America.

Smith et al.

121

AcknowledgmentsWe greatly appreciate Michael Kunz for helpful feedback re-garding the Mesa material, as well as his and Bill Hedman’s assistance in accessing the Putu and Mesa collections at the Bureau of Land Management, Fairbanks, AK. We owe many thanks to Dennis Stanford, Pegi Jodry, and Joe Gingerich for providing invaluable assistance in accessing the Alaskan fluted point and Irwin Sluiceway collections at the National Museum of Natural History, Washington, D.C. Thanks also to Jim Whit-ney, Sam Coffman, and Scott Shirar of the University of Alaska Museum of the North, Fairbanks, AK, for providing access to fluted Alaskan collections curated at the museum. Special thanks to Bob Gal and Jane Lakeman, who provided access to Alaskan fluted points curated at National Park Service offices in Anchorage, AK. We’d like to acknowledge Bob Gal, Dennis Stanford, Mike Kunz, and Paul Matheus, who generously shared ideas or unpublished data, but shouldn’t be assumed to agree with the interpretations presented here. Finally, thanks to Don Dumond and an anonymous reviewer, who returned helpful comments on an earlier draft of this paper. Our research on Northern Paleoindian technologies has been supported by the National Park Service (Noatak National Preserve), the National Park Service Shared Beringia Heritage Program, and the fol-lowing National Science Foundation grants: Goebel (1019190), Rasic (0227962 and 1111395), and Smith (1204085).

References citedAlexander, H. L. 1987 Putu: A Fluted Point Site in Alaska. Depart-ment of Archaeology, Publication No. 17. Simon Fraser University, Burnaby, British Columbia.

Anderson, D. D. 1972 An archaeological survey of Noatak Drain-age, Alaska. Arctic Anthropology 9(1):66–117.

Bamforth, D. B. 1991 Flintknapping skill, communal hunting, and Paleoindian projectile point typology. Plains Anthropologist 36(137):309–22.

Bever, M. R. 2000 Paleoindian Lithic Technology and Landscape Use in Late Pleistocene Alaska: A Study of the Mesa Complex. Un-published Ph.D. dissertation, Department of Anthropology, South-ern Methodist University, Dallas, Texas.

——— 2001 An overview of Alaskan late Pleistocene archaeol-ogy: Historical themes and current perspectives. Journal of World Prehistory 15(2):125–91.

——— 2006a Rethinking the Putu site: Results of a spatial analysis of a fluted point site in northern Alaska. Arctic Anthropology 43(1):20–39.

——— 2006b Too little, too late? The radiocarbon chronology of Alaska and the peopling of the New World. American Antiquity 71(4):595–620.

——— 2008 Distinguishing Holocene microblades from a Paleo-indian component at the Mesa site, Alaska. Journal of Field Archaeol-ogy 33:133–50.

Binford, L. R. 1980 Willow smoke and dogs’ tails: Hunter-gath-erer settlement systems and archaeological site formation. American Antiquity 45(1):4–20.

Bryan, A. L. 1969 Early man in America and the late Pleistocene chronology of western Canada and Alaska. Current Anthropology 10(4):339–65.

Bronk, R. C. 2009 Bayesian analysis of radiocarbon dates. Radio-carbon 51(1):337–60.

Churchill, S. 1993 Weapon technology, prey size selection, and hunting methods in modern hunter-gatherers: Implications for hunting in the Palaeolithic and Mesolithic. Archeological Papers of the American Anthropological Association 4(1):11–14.

Cinq-Mars, J., C. R. Harington, D. E. Nelson, and R. S. MacNeish 1991 Engigstciak revisited: A note on early Holocene AMS dates from the “Buffalo Pit.” In NOGAP Archaeology Project: An Integrated Archaeological Research and Management Approach, edited by J. Cinq-Mars and J. L. Pilon, pp. 33–44. Occasional Paper Number 1, Cana-dian Archaeological Association, Montreal.

Clark, D. W. 1984 Northern fluted points: Paleo-Eskimo, Paleo-Arctic, or Paleo-Indian. Canadian Journal of Anthropology 4(1):65–81.

——— 1991 The northern (Alaska-Yukon) fluted points. In Clovis Origins and Adaptation, edited by R. Bonnichsen and K. L. Turnmire. pp. 35–48. Center for the Study of the First Americans, Oregon State University Press, Corvallis.

——— 2009 A reconsideration of the British mountain assem-blage from the Engigstciak site, northern Yukon. In Painting the Past with a Broad Brush: Papers in Honour of James Valliere Wright, edited by D. Keelyside and J.-L. Pilon, pp. 515–27. Papers of the Canadian Museum of Civilization, Gatineau, Quebec.

Clark, D. W., and A. M. Clark 1983 Paleo-Indians and fluted points: Subarctic alternatives. Plains Anthropology 28:283–92.

——— 1993 Batza Téna: The Trail to Obsidian. Archaeological Survey of Canada Mercury Series Paper 147. Canadian Museum of Civilization, Hull.

Cook, J. P. 1969 The Early Prehistory of Healy Lake, Alaska. Unpub-lished Ph.D. dissertation, Department of Anthropology, University of Wisconsin, Madison.

——— 1995 Characterization and distribution of obsidian in Alaska. Arctic Anthropology 32(1):92–100.

Dixon, E. J. 1985 Cultural chronology of central interior Alaska. Arctic Anthropology 22(1):47–66.

——— 1999 Bones, Boats and Bison: Archaeology and the First Colo-nization of Western North America. University of New Mexico Press, Albuquerque.

——— 2001 Human colonization of the Americas: Timing, tech-nology and process. Quaternary Science Reviews 20:277–99.

——— 2011 Arrows, atlatls, and cultural-historical conundrums. In From the Yenisei to the Yukon: Interpreting Lithic Assemblage Variability in Late Pleistocene/Early Holocene Beringia, edited by T. Goebel and I. Buvit, pp. 362–70. Texas A&M University Press, College Station.

Dumond, D. E. 1980 The archaeology of Alaska and the peopling of America. Science 209:984–91.

——— 2001 The archaeology of eastern Beringia: Some con-trasts and connections. Arctic Anthropology 38(2):196–205.

Eerkens, J. W., and C. P. Lipo 2005 Cultural transmission, copying errors, and the generation of variation in material culture and the ar-chaeological record. Journal of Anthropological Archaeology 24:316–34.

Ellis, C. J., D. H. Carr, and T. J. Loebel 2011 The Younger Dryas and late Pleistocene peoples of the Great Lakes region. Quaternary International 242:534–45.

Fiedel, S. J. 2011 The mysterious onset of the Younger Dryas. Quaternary International 242(2):262–66.

Gal, R. 1976 Paleo-Indians of the Brooks Range: A tradition of

Biface Traditions of Northern Alaska and Their Role in the Peopling of the Americas

122

uncontrolled comparison. Paper Presented at the 41st Annual Meet-ing of the Society for American Archaeology, St. Louis.

——— 1982 Excavation of the Tunalik site, Northwestern National Petroleum Reserve in Alaska. Anthropological Papers of the University of Alaska 20(1-2):61–78.

Goebel, T., and B. A. Potter in press First traces: Late Pleistocene human settlement of the Arctic. In Handbook of Arctic Archaeology, edited by O. K. Mason and M. T. Friesen. Cambridge University Press, Cambridge.

Goebel, T., and S. B. Slobodin 1999 The colonization of western Beringia: Technology, ecology, and adaptations. In Ice Age Peoples of North America: Environments, Origins, and Adaptations, edited by R. Bonnichsen and K. L. Turnmire, pp. 104–55. Center for the Study of the First Americans, Oregon State University, Corvallis.

Goebel, T., R. Powers, and N. Bigelow 1991 The Nenana complex of Alaska and Clovis origins. In Clovis Origins and Adaptations, edited by R. Bonnichsen and K. L. Turnmire, pp. 49–79. Center for the Study of the First Americans, Oregon State University, Corvallis.

Goebel, T., H. L. Smith, L. DiPietro, M. R. Waters, B. Hockett, K. E. Graf, R. Gal, S. B. Slobodin, R. J. Speakman, S. G. Driese, and D. Rhode 2013 Serpentine Hot Springs, Alaska: Results of excava-tions and implications for the age and significance of northern flut-ed points. Journal of Archaeological Science 40(12):4222–33.

Hamilton, T. D., and T. Goebel 1999 Late Pleistocene peopling of Alaska. In Ice Age Peoples of North America: Environments, Origins, and Adaptations, edited by R. Bonnichsen and K. L. Turnmire, pp. 156–99. Center for the Study of the First Americans, Oregon State University Press, Corvallis.

Haynes, C. V., Jr. 1987 Clovis Origin Update. Kiva (52)2:83–93.

Hedman, B. 2010 The Raven Bluff Site: Preliminary Findings from a Late Pleistocene Site in the Alaskan Arctic. U.S. Department of the Interior, Bureau of Land Management, 7.

Hoffecker, J. F. 2001 Late Pleistocene and early Holocene sites in the Nenana River Valley, central Alaska. Arctic Anthropology 38(2):139–53.

——— 2005 Incredible journey: Plains bison hunters in the Arc-tic (Review of Kunz, Bever, and Adkins 2003). Review of Archaeology 26(2):18–22.

——— 2011 Assemblage variability in Beringia: the Mesa factor. In From the Yenisei to the Yukon: Interpreting Lithic Assemblage Variability in Late Pleistocene/Early Holocene Beringia, edited by T. Goebel and I. Buvit, pp. 165–78. Texas A&M University Press, College Station.

Hoffecker, J. F., and S. A. Elias 2007 Human Ecology of Beringia. Columbia University Press, New York.

Hoffecker, J. F., W. R. Powers, and T. Goebel 1993 The colonization of Beringia and the peopling of the New World. Science 259:46–53.

Holmes, C. E. 2001 Tanana River valley archaeology circa 14,000 to 9000 B.P. Arctic Anthropology 38(2):154–70.

Holmes, C. E., B. A. Potter, J. D. Reither, O. K. Mason, R. M. Thorson, and P. M. Bowers 2008 Geological and cultural context of the Nogahabara I site. American Antiquity 73(4):781–90.

Kelly, R., and L. Todd 1988 Coming into the country. American Antiquity 53(2):231–44.

King, M. L., and S. B. Slobodin 1996 A fluted point from the Up-tar site, northeastern Siberia. Science 273:634–36.

Kunz, M. L. 1982 The Mesa site: An Early Holocene Hunting Stand in the Iteriak Valley, Northern Alaska. Anthropological Papers of the University of Alaska 20(1-2):113–22.

Kunz, M. L., and R. E. Reanier 1994 Paleoindians in Beringia: Evi-dence from arctic Alaska. Science 263:660–62.

——— 1995 The Mesa site: A Paleoindian hunting lookout in arctic Alaska. Arctic Anthropology 32(1):5–30.

Kunz, M., M. Bever, and C. Adkins 2003 The Mesa site: Paleoindians above the arctic Circle. BLM-Alaska Open File Report 86, U.S. Depart-ment of the Interior, Bureau of Land Management, Anchorage.

Loy, T. H., and E. J. Dixon 1998 Blood residues on fluted points from eastern Beringia. American Antiquity 63(1):21–46.

MacNeish, R. S. 1956 The Engigstciak Site on the Yukon Arctic Coast. Anthropological Papers of the University of Alaska 4:91–111.

——— 1959 Men out of Asia: As Seen from the Northwest Yukon. Anthropological Papers of the University of Alaska 7(2):41–61.

Mackay, J. R., W. H. Mathews, and R. S. MacNeish 1961 Geol-ogy of the Engigstciak archaeological site, Yukon Territory. Arctic 14(1):25–52.

Malyk-Selivanova, N., G. M. Ashley, R. Gal, M. D. Glascock, and H. Neff 1998 Geological-geochemical approach to “sourcing” of prehistoric chert artifacts, northwestern Alaska. Geoarchaeology 13(7):673–708.

Mann, L. H., R. F. Reanier, D. M. Peteet, M. L. Kunz, and M. John-son 2001 Environmental change and arctic Paleoindians. Arctic Anthropology 38(2):119–38.

Mann, D. H., P. Groves, M. L. Kunz, R. E. Reanier, and B. V. Gaglioti 2013 Ice-age megafauna in arctic Alaska: Extinction, invasion, sur-vival. Quaternary Science Reviews 70:91–108.

Meltzer, D. J. 1993 Pleistocene peopling of the Americas. Evolu-tionary Anthropology 1:157–69.

——— 2009 First Peoples in a New World: Colonizing Ice Age America. University of California Press, Berkeley.

Morlan, R. E. 2003 Current perspectives on the Pleistocene ar-chaeology of eastern Beringia. Quaternary Research 60:123–32.

Mull, C. G. 1995 The Geological Distribution of Chert in the Brooks Range. State of Alaska Department of Natural Resources, Division of Geological and Geophysical Surveys, Public-Data File 95-32, Fair-banks.

Odess, D., and J. T. Rasic 2007 Toolkit composition and assem-blage variability: The implications of Nogahabara I, northern Alaska. American Antiquity 72(4):691–717.

Potter, B. A. 2008 Radiocarbon and chronology of central Alaska: The technological continuity and economic change. Radiocarbon 50(2):181–204.

——— 2011 Late Pleistocene and early Holocene assemblage variability in central Alaska. In From the Yenisei to the Yukon: Inter-preting Lithic Assemblage Variability in Late Pleistocene/Early Holocene Beringia, edited by T. Goebel and I. Buvit pp. 215–33. Texas A&M University Press, College Station.

Powers, W. R., and J. F. Hoffecker 1989 Late Pleistocene set-tlement in the Nenana Valley, central Alaska. American Antiquity 54(2):263–87.

Rasic, J. 2008 PaleoAlaskan Adaptive Strategies Viewed from Northwest Alaska. Unpublished Ph.D. dissertation, Department of Anthropology, Washington State University, Pullman, Washington.

——— 2011 Functional variability in the late Pleistocene ar-chaeological record of eastern Beringia: A model of late Pleistocene land use and technology from northwest Alaska. In From the Yenisei to the Yukon: Interpreting Lithic Assemblage Variability in Late Pleistocene/

Smith et al.

123

Early Holocene Beringia, edited by T. Goebel and I. Buvit, pp.1–30. Texas A&M University Press, College Station.

Rasic, J. and P. E. Matheus 2007 A reconsideration of purported Holocene bison in northern Alaska. Arctic 60(4):381–86.

Reanier, R. E. 1994 The Putu site: Pleistocene or Holocene? Cur-rent Research in the Pleistocene11:148–51.

——— 1995 The antiquity of Paleoindian materials in northern Alaska. Arctic Anthropology 32:31–50.

Reimer, P. J., M. G. L. Baillie, E. Bard, A. Bayliss, J. W. Beck, P. G. Black-well, C. B. Ramsey, C. E. Buck, G. S. Burr, R. L. Edwards, M. Friedrich, P. M. Grootes, T. P. Guilderson, I. Hajdas, T. J. Heaton, A. G. Hogg, K. A. Hughen, K. F. Kaiser, B. Kromer, F. G. McCormac, S. W. Manning, R. W. Reimer, D. A. Richards, J. R. Southon, S. Talamo, C. S. M. Tur-ney, J. van der Plicht, and C. E. Weyhenmeyer 2009 IntCal09 and Marine09 radiocarbon age calibration curves, 0–50,000 years cal yr BP. Radiocarbon 51(4):1111–50.

Saleeby, B. 2010 Ancient footsteps in a new land: Building an in-ventory of the earliest Alaskan sites. Arctic Anthropology 47(2):116–32.

Sellet, F. 2001 A changing perspective on Paleoindian chronol-ogy and typology: A view from the northwestern Plains. Arctic An-thropology 38(2):48–63.

Shapiro, B., A. J. Drummond, A. Rambaut, M. C. Wilson, P. E. Matheus, A. V. Sher, O. G. Pybus, M. T. P. Gilbert, I. Barnes, J. Binladen, E. Will-erslev, A. J. Hansen, G. F. Baryshnikov, J. A. Burns, S. Davydov, J. C. Driver, D. G. Froese, C. R. Harington, G. Keddie, P. Kosintsev, M. L. Kunz, L. D. Martin, R. O. Stephenson, J. Storer, R. Tedford, S. Zimov, and A. Cooper 2004 Rise and fall of the Beringian steppe bison. Science 306:1561–65.

Slobodin, S. 2001 Western Beringia at the end of the Ice Age. Arctic Anthropology 38(2):31–47.

——— 2011 Late Pleistocene and early Holocene cultures of Beringia: The general and the specific. In From the Yenisei to the Yu-kon: Interpreting Lithic Assemblage Variability in Late Pleistocene/Early Holocene Beringia, edited by T. Goebel and I. Buvit, pp.91–116. Texas A&M University Press, College Station.

Torrence, R. 2002 Thinking Big about Small Tools. Archeological Papers of the American Anthropological Association 12(1):179–89.

West, F. H. 1996 The archaeological evidence. In American Begin-nings: The Prehistory and Paleoecology of Beringia, edited by F. H. West, pp. 537–60. University of Chicago Press, Chicago.

Willey, G. R. 1966 An Introduction to American Archaeology. North and Middle America, vol. 1. Prentice Hall, Englewood Cliffs, New Jersey.

Wilson, M. C., L. V. Hills, and B. Shapiro 2008 Late Pleistocene northward-dispersing Bison antiquus from the Bighill Creek forma-tion, Gallelli Gravel Pit, Alberta, Canada, and the fate of Bison occiden-talis. Canadian Journal of Earth Sciences 45:827–59.

Wormington, H. M. 1957 Ancient Man in North America, fourth ed. Popular Series No. 4. Denver Museum of Natural History, Denver.

Young, C., and S. Gilbert-Young 2007 A fluted projectile-point base from Bering Land Bridge National Preserve, northwest Alaska. Current Research in the Pleistocene 24:154–56.

Biface Traditions of Northern Alaska and Their Role in the Peopling of the Americas