fulltext (2)

49
A Tale of Two Migrations: Reconciling Recent Biological and Archaeological Evidence for the Pleistocene Peopling of the Americas Bonnie L. Pitblado Published online: 12 March 2011 Ó Springer Science+Business Media, LLC 2011 Abstract This article synthesizes the 2000s-era ‘‘peopling of the Americas’’ data drawn from molecular biology, osteology, and archaeology. Collectively, they suggest that colonization proceeded in two pulses, both originating in western Beringia, and before that, south-central and southeastern Siberia. The first pulse occurred circa 16 k–15 k cal. B.P. by watercraft along the coast of Beringia and western North and South America. The second took place 1,000 years later and involved proto-Clovis hunter-gatherers who used the ice-free corridor as a conduit south. At least eight North American sites dating as far back as the Last Glacial Maximum suggest that the peopling picture may eventually need to change to accommodate an earlier than previously thought migration through the ice-free corridor. For now, the data are not strong enough to support this scenario, but they are tantalizingly close. Keywords Peopling Á Colonization Á Migration Á Americas Á New World Introduction In 2000, in the pages of this journal, Fiedel overviewed then-current evidence, theories, and future research directions in the always controversial and fertile ‘‘peopling of the New World’’ arena. I pick up where Fiedel left off, summarizing how far we have come in the first decade of the 21st century in elucidating the who, when, where, and how of New World colonization. I do not reiterate ground covered by Fiedel. He did a comprehensive job of bringing readers up to turn-of-the- millennium speed. In addition, the peopling literature is expanding at an exponential rate as geneticists, biologists, oceanographers, and other earth, life, and even B. L. Pitblado (&) Anthropology Program, Utah State University, 0730 Old Main Hill, Logan, UT 84322-0730, USA e-mail: [email protected] 123 J Archaeol Res (2011) 19:327–375 DOI 10.1007/s10814-011-9049-y

Upload: mikeytipsword

Post on 27-Oct-2014

52 views

Category:

Documents


11 download

TRANSCRIPT

Page 1: fulltext (2)

A Tale of Two Migrations: Reconciling RecentBiological and Archaeological Evidencefor the Pleistocene Peopling of the Americas

Bonnie L. Pitblado

Published online: 12 March 2011

� Springer Science+Business Media, LLC 2011

Abstract This article synthesizes the 2000s-era ‘‘peopling of the Americas’’ data

drawn from molecular biology, osteology, and archaeology. Collectively, they

suggest that colonization proceeded in two pulses, both originating in western

Beringia, and before that, south-central and southeastern Siberia. The first pulse

occurred circa 16 k–15 k cal. B.P. by watercraft along the coast of Beringia and

western North and South America. The second took place 1,000 years later and

involved proto-Clovis hunter-gatherers who used the ice-free corridor as a conduit

south. At least eight North American sites dating as far back as the Last Glacial

Maximum suggest that the peopling picture may eventually need to change to

accommodate an earlier than previously thought migration through the ice-free

corridor. For now, the data are not strong enough to support this scenario, but they

are tantalizingly close.

Keywords Peopling � Colonization � Migration � Americas � New World

Introduction

In 2000, in the pages of this journal, Fiedel overviewed then-current evidence,

theories, and future research directions in the always controversial and fertile

‘‘peopling of the New World’’ arena. I pick up where Fiedel left off, summarizing

how far we have come in the first decade of the 21st century in elucidating the who,

when, where, and how of New World colonization. I do not reiterate ground covered

by Fiedel. He did a comprehensive job of bringing readers up to turn-of-the-

millennium speed. In addition, the peopling literature is expanding at an exponential

rate as geneticists, biologists, oceanographers, and other earth, life, and even

B. L. Pitblado (&)

Anthropology Program, Utah State University, 0730 Old Main Hill, Logan, UT 84322-0730, USA

e-mail: [email protected]

123

J Archaeol Res (2011) 19:327–375

DOI 10.1007/s10814-011-9049-y

Page 2: fulltext (2)

planetary scientists join archaeologists in the ongoing effort to illuminate

colonization of the Americas. Summarizing developments of the last decade is a

sufficiently daunting challenge without harkening back to Folsom, New Mexico, in

the roaring 1920s.

Fiedel (2000, pp. 84-86) concluded his piece by identifying future research

avenues that he viewed as likely to bear fruit: expanding genetic and residue studies;

refining absolute dating techniques, including radiocarbon and optically stimulated

luminescence (OSL); excavating more stratified Late Pleistocene sites; conducting

new research in Mongolia, northern China, Siberia, South America, and near the

ice-free corridor (IFC); and using simulation models to illuminate extinction and

colonization processes. Many of these research directions have been vigorously

pursued with thought-provoking results. Other research agendas and finds of the last

decade fall more into the ‘‘wow—I never saw that coming’’ class: 14,300–14,100-

year-old human coprolites as smoking guns of pre-Clovis presence in North

America come immediately to mind.

I preface my peopling overview with an observation about the mindset of

contemporary participants in the dialogue. Most who weigh in on various peopling

debates lament how many unanswered questions remain and how little we really

know about the initial human colonization of the Americas. I myself began this

synthetic enterprise less than optimistic that I would better understand the peopling

process after typing my last sentence. I was therefore pleased to find myself

concluding, after reading over 300 manuscripts from the 2000s and reviewing many

older classics, that when the evidence from myriad disciplines is viewed in concert

per Dillehay’s (2009) recent call to action, there are more answers embedded

therein—more tangible constraints on what colonizing humans could and could not

have done—than I anticipated.

The root problem is not so much that we lack the data needed to draw well-

reasoned conclusions on controversial peopling issues. Instead, it is that few

archaeologists have either the time to keep up with the volume of peopling-related

work published in journals ranging from American Journal of Human Genetics to

Earth and Planetary Science Letters or the expertise to critically assess data that

appear in those wide-ranging outlets. Without advanced training in molecular

biology, how does the average archaeological Jane evaluate the relative merits of

sequencing the D-loop versus the entire mtDNA genome to detect variation within

or among human populations? Similarly, when 26 Proceedings of the NationalAcademy of Sciences co-authors (Firestone et al. 2007) argue that microspherules

and nanodiamonds indicate that the Younger Dryas began with an extraterrestrial

bang powerful enough to have impacted Clovis demography (not a topic I discuss

here, but a hot one), how does even the most widely read archaeologist begin to

weigh the evidence proposed to support the contention?

That said, I marshal as much evidence as possible and ascribe more weight to

those data or approaches that are most convincing to me and to the scientists

working within a given discipline, and to conclusions that approach consensus

support. As Yesner et al. (2004, p. 200) noted, ‘‘The goal of science is to create the

best narratives that we can from the evidence at hand, but to be willing to rewrite

those narratives in an instant if and when superior evidence becomes available.’’

328 J Archaeol Res (2011) 19:327–375

123

Page 3: fulltext (2)

I have no doubt that circa 2020, after hundreds more studies have appeared, my

narrative will require revision. I also conclude that we have the tools now to develop

well-supported peopling scenarios that can be presented without undue hand-

wringing over questions that remain unresolved. The scenarios are not perfect, but

peopling scholars are doing an admirable job of garnering relevant evidence.

Here I argue that recent data converge to suggest that the peopling of the New

World occurred in two pulses, both originating in southern Siberia. The first

proceeded along the Pacific Rim and coast of Alaska via watercraft

16,000–15,000 years ago, and the second 1,000 years later on land by way of

Beringia and the ice-free corridor. I build my case by first commenting on the

‘‘traditional’’ peopling model and the impact it has had on peopling studies. I then

offer a list of findings I see as so well supported that any peopling scenario must

account for them. I move next to a discussion of the contributions molecular

biologists have made recently to our understanding of the origins of the First

Americans, the timing of initial peopling, and likely numbers of founding

populations and individuals constituting those populations. I then overview recently

gathered osteological and archaeological evidence derived from northeast Asia,

Beringia, and the Americas. That evidence does not include a detailed discussion of

lithic technology; I agree with Waguespack (2007, p. 66), who noted that ‘‘we are

no closer to solving the problem of tracing migration through prehistoric material

culture than we were 80 or even 150 years ago.’’ I end by pointing out weaknesses

in my scenario, an alternative scenario that could account for those weaknesses, and

future research directions peopling scholars might most profitably pursue.

The traditional ‘‘peopling’’ perspective

Innumerable ‘‘peopling of the New World’’ manuscripts begin with a statement to

this effect: ‘‘The traditional model of the peopling of the New World holds that

Clovis ancestors crossed the Bering Land Bridge and moved south through the ice-

free corridor’’ (e.g., Bryan and Gruhn 2003; Davis et al. 2002-2004; Dillehay 2009;

Dixon 2001; Erickson et al. 2005; Fagundes et al. 2008; Faught 2008; Mandryk

et al. 2001). The statement is often offered without citations, however, presumably

because it is viewed as so obviously true. But is it true? If one seeks citations that

demonstrate long-term, wholesale support for the Clovis First-Beringia-IFC

peopling model, one fails to find them. Acosta’s (2002 [1604]) natural history of

the Indies pops up in reference lists, as does a 45-year-old Science piece by Haynes

(1964). A few writers have more recently expressed support for all or parts of the

model (e.g., Fiedel 1999, 2002, 2004; Goebel 1999; Haynes 2002). But, in fact, the

number of archaeologists who claim that there is a ‘‘traditional’’ peopling model far

exceeds the number of archaeologists who have ever advocated the model in print.

That said, the mere perception that a predominant model has infiltrated our

collective psyche has profoundly shaped the way archaeologists have approached

peopling problems, and not in a good way. Many past efforts to advance peopling

debates have amounted to the presentation of ‘‘site X’’ as demonstrating that

‘‘Clovis First’’ is wrong (with Clovis defined recently by Waters and Stafford

J Archaeol Res (2011) 19:327–375 329

123

Page 4: fulltext (2)

[2007a] as dating from 13,250 to 12,800 cal. B.P.). These efforts include frenzied

scrutiny of the site that reveals flaws of varying magnitude but always sufficient to

induce doubt that ‘‘site X’’ is a paradigm buster. And that is the problem. Our

understanding of the peopling of the New World should not rest on the shoulders of

a single site, because every locality in the world, regardless of age or location, is at

best a nature-ravaged snapshot of one moment in prehistory, excavated by imperfect

human beings. How could there not be flaws?

I suggest that those attempting to advance understanding of the peopling of the

New World should instead, returning to Yesner et al.’s (2004) modest suggestion,

create the best narratives possible given the evidence available, or, as Turner (2002)

put it, develop a peopling hypothesis that is maximally holistic, parsimonious, and

offers the greatest concordance of disparate lines of evidence. We should endeavor

to reconcile as many well-established data points as we can, being explicit about

what we consider solid evidence and where we perceive gaps that could guide future

research or that might even be deal-breakers for our own scenarios. Some have done

this (e.g., Bradley and Stanford 2004; Goebel et al. 2008). If more of us adopt this

approach, rather than erecting artificial ‘‘bars,’’ Clovis or otherwise, we will

illuminate issues related to the peopling of the New World much more quickly and,

one hopes, more civilly than has often been the case for the last century.

In this spirit, I introduce my narrative with a list of findings that I think are

sufficiently well established and important that a credible peopling scenario must

account for them. (1) A number of sites in North and South America have been

convincingly shown to predate Clovis, but none of the most convincing or widely

accepted more than slightly predates c. 15 k cal. B.P. We can debate which sites

make the list, but even the most skeptical among us must recognize that such a list

exists. (2) Some sites unequivocally shown to date to Clovis time do not share the

Clovis predilection for mammoth hunting (e.g., marine-oriented Quebrada Jaguay,

Peru). (3) Genetically, Native Americans today express a limited number of mtDNA

and NRY haplogroups, but most are widespread in the hemisphere. They share that

range of haplogroups with precious few Old World populations. First Nations

people also share a few nuclear DNA markers with populations in western Beringia

and nowhere else. (4) The oldest New World skeletal remains show a generalized

morphology akin to that of Australians and Africans but different from Native

Americans and East Asians. (5) The Pacific coast of Beringia and the Northwest was

ice free-by 16,000-15,000 years ago. A corridor between the Laurentide and

Cordilleran Ice Sheets deglaciated later, c. 14,000-13,500 years ago.

The scenario that for me best reconciles these ‘‘anchor points’’ shares many

elements in common with models proposed recently by Dixon (1999, 2001),

Erlandson (2002), Fagundes et al. (2008), Kitchen et al. (2008), Goebel et al. (2008),

and most especially Mandryk et al. (2001). I overview the scenario, then discuss and

offer evidence for its central components, beginning with genetic data and then

moving through osteological and archaeological findings.

The First Americans immigrated to the New World c. 16 k–15 k cal. B.P.,

probably at the later end of that range. They initiated their odyssey in south-central

and southeastern Siberia before the Last Glacial Maximum (LGM) of 24,000-

21,000 years ago, moving north in Siberia along major rivers, including the Yenesei

330 J Archaeol Res (2011) 19:327–375

123

Page 5: fulltext (2)

and Lena, then east to western Beringia, where populations paused before, during,

and immediately after the LGM. The Amur River Basin/Sea of Okhotsk in

southeastern Siberia also served as a point of origin for ancestors of First Nations

people, with groups moving north along the Amur River, then farther north along or

across the Sea of Okhotsk to Beringia, where they, too, paused perhaps during and

certainly just after the LGM.

People entered the New World in two recognizable pulses (and possibly many

more that we cannot and will never ‘‘see’’). The first immigrants employed boats to

follow the coastline of Alaska and the Pacific Northwest, which was ice-free by

16,000-15,000 years ago. Those founders initially retained a focus on coastal,

riverine, and lacustrine resources as descendants made their way inland to southern

Oregon’s Pluvial Lake Chewuacan and south to Chile’s Monte Verde by 14,600 cal.

B.P. The second group of immigrants traversed Beringia and made their way south

via an IFC sometime shortly after waning glacial conditions exposed it 14,000-

13,500 years ago. Adapted to terrestrial hunting, their descendants moved east,

west, and south, refining their hunting technology and soon perfecting the Clovis

tool kit that has been documented in so many contexts across North and Central

America.

Geographic origins of the First Americans from a genetic perspective

Evidence placing the geographic origins of the First Americans in south-central and

southeastern Siberia includes a staggering array of genetic studies encompassing

mtDNA, NRY, and nuclear DNA. For accessible overviews of molecular

anthropology, see Merriwether (2006), Meltzer (2009), and O’Rourke et al.

(2000). Of the many dozen peopling-specific investigations conducted in the past

decade, none suggests an immediate point of origin for the First Americans other

than northeast Asia, and a significant number specify the particular regions cited in

my scenario. Even so, a cursory reading of relevant genetic literature might suggest

that there is significant disagreement among molecular scientists about the location

of the Old World homeland of the First Americans. In fact, they do not so much

disagree about where New World colonists began their journey as they vary in the

level of geographic specificity to which they can or will commit. In addition,

depending on the goals and methods of a particular genetic study, resultant data can

reveal different staging points in the peopling process, and sometimes more than

one staging area. We can definitively rule out, after all, the notion that founding

populations spontaneously generated in south-central Siberia or anywhere else in

Asia, and indeed, some have tracked the origins of First Americans farther back in

time and farther south and west in space than their final ‘‘jumping off’’ spot (e.g.,

Comas et al. 2004; Goebel 2007).

To illustrate my semantic point: various contemporary authors have concluded

from their genetic data sets that the likeliest point of origin for the peopling of the New

World is southern Siberia (Derenko et al. 2007; Wells et al. 2001), or south-central

Siberia (Bortolini et al. 2003; Schurr 2004; Schurr and Sherry 2004), or central Siberia

(Lell et al. 2002), or eastern Siberia (Wang et al. 2007), or northern Siberia (Forster

J Archaeol Res (2011) 19:327–375 331

123

Page 6: fulltext (2)

2004), or just plain old Siberia (Seielstad et al. 2003). Others point to Asia (O’Rourke

2009), or East Asia (Bandelt et al. 2003; Malhi et al. 2007), or northeast Asia

(Battilana et al. 2006; Fagundes et al. 2008), or Mongolia (Malhi et al. 2002), or the

Altai Mountains (Dornelles et al. 2005; Malhi and Smith 2002; Santos et al. 1999,

Starikovskaya et al. 2005; Zegura et al. 2004), or the ‘‘area around Lake Baikal’’

(Derenko et al. 2001; Eshleman et al. 2003). Yet others implicate the Lower Amur/Sea

of Okhotsk region (Lell et al. 2002; Starikovskaya et al. 2005) or western Beringia

(Derbeneva et al. 2002; Schroeder et al. 2007, 2009; Tamm et al. 2007).

This would appear to be a long list of possible source regions for the First

Americans. Yet these varying geographic attributions overlap each other to the point

that all of them can be reconciled in the peopling scenario I outlined, which

identified south-central Siberia, southeastern Siberia, and western Beringia as

staging points for immigration to the New World (Fig. 1). All three of those regions

are encompassed by the broader designations Siberia, central Siberia, East Asia,

northeast Asia, and Asia. The Altai Mountains and other iterations of the Altai

region and Lake Baikal are located in south-central Siberia, and Mongolia abuts

them to the south, separated from them by a modern political border. Southeastern

Siberia comprises the Lower Amur and Sea of Okhotsk regions; northern Siberia

and Kamchatka encompass western Beringia. What is striking is not how many

different potential points of origin genetic data suggest, but rather how few, and how

consistent they really are.

Fig. 1 Map depicting Asian regions pertinent to the peopling of the New World, together withlandforms, rivers, cities, and archaeological sites mentioned in the text. (1) Yana Rhino Horn Site; (2)Berelekh; (3) Ikhine 2; (4) Verkhne-Troitskaya; (5) Ushki Lake; (6) Mal’ta. The dotted line in thenortheast depicts the approximate extent of Beringia at the Last Glacial Maximum (LGM). Figure draftedby Holly Andrew

332 J Archaeol Res (2011) 19:327–375

123

Page 7: fulltext (2)

How robust are these genetic data? Very. Genetic data generated this century

include assessments of modern and ancient mtDNA (found in cell cytoplasm and

passed from mothers to offspring), the nonrecombining portion of the Y

chromosome (NRY) (passed from fathers to sons) (Bortolini et al. 2003; Eshleman

et al. 2003; Lell et al. 2002; Santos et al. 1999; Seielstad et al. 2003), and nuclear

(autosomal) DNA. The most common studies, those focusing on mtDNA, invoke

different and in some studies multiple methodologies, including sequencing of the

mutation-prone D-loop (Derenko et al. 2001; Starikovskaya et al. 2005), sequencing

base-pair segments outside the D-loop (Silva et al. 2002), and sequencing complete

or nearly complete mtDNA (Bandelt et al. 2003; Derenko et al. 2007; Fagundes

et al. 2008; Gilbert et al. 2008a; Macaulay et al. 2005; Reidla et al. 2003;

Starikovskaya et al. 2005). Geneticists researching nuclear DNA focus on a variety

of mutations present or not present in populations in the New and Old Worlds

(Battilana et al. 2006, 2007; Schroeder et al. 2007, 2009; Wang et al. 2007). Many

molecular biologists extract mtDNA and NRY data sets from the same populations

(Schurr 2004; Schurr and Sherry 2004) or ancient human remains (Kemp et al.

2007) to independently evaluate maternal and paternal demographic histories.

In specific terms, the genetic studies that point to south-central and southeastern

Asia as geographic source areas for New World founders note that these two

regions, and no others, are now home to populations that share an array of

haplogroups in common with Native Americans. (A haplogroup is a grouping of

similar and ancestrally related haplotypes; a haplotype represents a set of alleles on

a single chromosome that are inherited together and thus shared by closely related

people.) Indigenous North and South American populations so far sampled

collectively exhibit seven mtDNA haplogroups (A, B, C, two variants of D, M,

and X) (e.g., Bandelt et al. 2003; Derenko et al. 2001; Fix 2005; Malhi et al. 2002,

2007; Perego et al. 2009; Schurr 2004; Schurr and Sherry 2004; Starikovskaya et al.

2005) and four NRY haplogroups (C, P, Q, and R) (e.g., Lell et al. 2002; Santos

et al. 1999; Schurr 2004; Seielstad et al. 2003; Tarazona-Santos and Santos 2002;

Zegura et al. 2004). Only peoples of the Altai region and the regions southeast and

southwest of Lake Baikal express a range of mtDNA haplogroups and NRY

haplogroups that encompasses those of First American populations (Derenko et al.

2001; Dornelles et al. 2005; Eshleman et al. 2003; Malhi et al. 2002, 2007; Malhi

and Smith 2002; Merriwether 2006; Schurr and Sherry 2004; Starikovskaya et al.

2005; Zegura et al. 2004).

Particular mtDNA and NRY haplotypes likewise suggest a link between First

Americans and Siberians of the Lower Amur/Sea of Okhotsk (mtDNA haplotypes

C1a in the Lower Amur and C1b among Native Americans, and NRY haplotypes

M45b and M173) (Lell et al. 2002; Starikovskaya et al. 2005). As Merriwether

(2006) has argued, it is more parsimonious to envision a single population

representing a half-dozen or so haplogroups colonizing the New World than it is to

envision multiple groups of immigrants each bearing one and only one haplogroup

arriving independently and then mixing until their collective haplogroup distribution

characterized descendant groups throughout the Americas. Silva et al. (2002) also

pointed out that similarity in nucleotide diversity in Native American haplogroups

J Archaeol Res (2011) 19:327–375 333

123

Page 8: fulltext (2)

suggests diversification occurred at the same time, indicating that all the

haplogroups share a common history.

Some genetic data also point to a moderate population bottleneck in western

Beringia, a place where migrants from the Altai and Amur regions paused for

perhaps thousands of years when ice sheets impeded access to the Americas (e.g.,

Kitchen et al. 2008; Mandryk et al. 2001). Studies of nuclear DNA indicate that

Native Americans show lower genetic diversity than populations on other continents

(Wang et al. 2007), an indicator of a bottleneck at some time, somewhere. The

presence and distribution of a particular mutation at locus D9S1120 (9RA) indicates

that the bottleneck occurred in western Beringia, because the mutation characterizes

just over 30% of DNA of all sampled North and South American Native American

populations and an identical percentage of Koryaks and Chukchis, the contemporary

residents of easternmost Siberia (Fig. 1). Elsewhere in Asia, including in the Altai

region, the percentage of populations showing this mutation is dramatically

different, indicating the mutation arose after movement of south-central Asians to

western Beringia. Bandelt et al. (2003), Derbeneva et al. (2002), Fagundes et al.

(2008), Forster (2004), Kitchen et al. (2008), and Tamm et al. (2007) also have

inferred a population bottleneck in western Beringia that ended at or shortly after

the LGM, basing their conclusions on mtDNA evidence. Santos et al. (1999)

showed that NRY data likewise indicate that people left south-central Siberia prior

to the LGM and lingered in western Beringia long enough for population

differentiation to occur.

Timing of the peopling of the Americas from a genetic perspective

The genetic data leave no room for doubt as to where the First Americans began

their voyage to the New World. Establishing the time frame for that migration has

always been more difficult. Telling time using molecular data requires adopting a

constant mutation rate for mtDNA or NRY to calculate how much time has elapsed

since two related populations diverged from each other. However, consensus on

those mutation rates has not been reached (e.g., Achilli et al. 2008; Armour et al.

1996; Fix 2005; Forster 2004; Forster et al. 2001; Ho and Endicott 2008; Ho and

Larson 2006; Kemp et al. 2007; Mishmar et al. 2003). Lack of consensus

notwithstanding, research of the past decade reflects not only methodological

refinement but also two interpretive shifts vis-a-vis the ground-breaking peopling-

related genetic work of the 1980–1990s. First, most researchers are deriving

significantly later dates for colonization of the New World than the 20,000-30,000

or more years ago often asserted in early studies (e.g., Torroni et al. 1992, 1993; see

Eshleman et al. 2003 for other examples). Second, their various estimates show

more consistency than they did in even the recent past, with most pointing to initial

colonization of the Americas shortly after the LGM.

In fact, close scrutiny of recent manuscripts inferring timing of initial peopling

from genetic data reveals a phenomenon similar to that noted for inferences

regarding geographic origins of First Americans. Those reporting particular dates

for colonization tend to place them in the rather tight and late range of

334 J Archaeol Res (2011) 19:327–375

123

Page 9: fulltext (2)

18 k-15 k cal. B.P. Some cite a colonization date significantly earlier or later than

this time frame, but with a very large standard deviation that encompasses 18,000-

15,000 years ago. Still others decline to cite particular dates, usually mentioning

problems with calibration of the molecular clock as their rationale, but offer

inferences consistent with a late colonization or inconsistent with an early one. And

some cite absolute dates for events that may appear to conflict with a late

colonization but that could actually be construed to support such a view. For

example, dates have been proposed for population divergences that occurred in Asia

well before colonization; i.e., the researchers established an upper limit for when

migration could have occurred, not the date it actually did.

Molecular biologists who have recently placed colonization of the New World in

the 18 k–15 k cal. B.P. range on the basis of mtDNA data include Atkinson et al.

(2008), Derbeneva et al. (2002), Fagundes et al. (2008), Gilbert et al. (2008a),

Kitchen et al. (2008), O’Rourke (2009), and Perego et al. (2009). Bortolini et al.

(2003), Seielstad et al. (2003), and Zegura et al. (2004) pinpointed the same time

frame for the initial peopling of the Americas using NRY data sets. Examples of

studies that offer date ranges that encompass this time frame include Kemp et al.’s

(2007) calculation that humans entered the Americas c. 13,500 years ago, but with a

95% confidence interval placing the event between 28,700 and 8,100 years ago.

Schurr (2004) and Schurr and Sherry (2004) similarly suggested colonization

occurred 20,000-15,000 years ago. Hey (2005), while acknowledging a low

probability that colonization could have occurred as many as 20,000 years ago,

demonstrated a much higher probability that it occurred in the past 15,000 years.

Eshleman et al. (2003) exemplify a research team that views peopling as having

occurred well after the 30 k cal. B.P. time frame inferred by many 1980s-1990s

studies but declines to assign an absolute date to the process based on the lack of

molecular clock consensus. Starikovskaya et al. (2005) similarly acknowledged

occupation of eastern Siberia by 30,000-28,000 years ago and view migration across

Beringia as occurring well after that but did not cite a particular time frame.

Number of founding population(s) and people from a genetic perspective

Two related subjects include the number of founding population(s) involved in the

peopling process and the number of people constituting those population(s). The

former has been the subject of more genetic research than the latter, but neither has

engendered consensus (Gruhn 2006). Still, with regard at least to the number of

migratory pulses that resulted in the peopling of the New World, most fall into one

of two camps that are not so very far removed from one another: that a single

population gave rise to all Pleistocene human occupation of the Americas, or that

two or slightly more contributed to the process. Few, if any, geneticists argue that

the peopling of the Americas occurred over an extended period of time and involved

the serial migration of many groups from multiple parts of Asia to the New World.

Those advocating the single-founder hypothesis include Fagundes et al. (2008),

Hey (2005), Kitchen et al. (2008), Malhi et al. (2002), Merriwether (2006),

and Silva et al. (2002), all of whom based their conclusions on mtDNA.

J Archaeol Res (2011) 19:327–375 335

123

Page 10: fulltext (2)

Bortolini et al. (2003) invoked NRY findings to conclude that a single male

migration contributed genetically to the peopling of the New World. Eshleman et al.

(2003), Santos et al. (1999), Tarazona-Santos and Santos (2002), and Zegura et al.

(2004) also interpreted their NRY data as supporting a single-founder scenario.

Schroeder et al. (2007, 2009) and Wang et al. (2007) drew the same conclusion from

autosomal DNA data that they argued reveal such consistency in the distribution of

highly distinctive alleles among contemporary Native Americans that only a single

founding population could have contributed genetically to them. Importantly, some

proponents of the single-founder model (e.g., Goebel et al. 2008; Hey 2005) have

noted that if the same source population generated a migratory pulse more than

once, this could have yielded descendant genetic signatures and distributions that

are essentially identical to one another, particularly given the gene flow that likely

took place among such populations before and after the Americas were settled.

In the multiple-founder camp, Bandelt et al. (2003), Derenko et al. (2007),

Forster (2004), O’Rourke (2009), Perego et al. (2009), Schurr (2004), Schurr and

Sherry (2004), Starikovskaya et al. (2005), and Tamm et al. (2007), drawing

inferences from their own and sometimes others’ mtDNA data, have argued that two

or perhaps more migrations spawned human occupation of the Americas. Perego

et al. (2009), for example, analyzed two rare mtDNA haplogroups in American and

Asian populations, concluding that only two independent migration events from

Asia can account for their current distribution in the Americas: one along the coast

of the Americas and one in northern North America. Lell et al. (2002), Schurr

(2004), and Schurr and Sherry (2004) drew the same conclusion based on NRY data.

Those arguing for multiple migrations from a genetic perspective generally cite,

as in the Perego et al. (2009) example, differential distributions of lineages in the

Americas as evidence for their position. Practitioners advocating the single-founder

model emphasize the opposite: the geographic ubiquity of a particular small suite of

haplogroups in the Americas that can all be traced to northeast Asia. These represent

different readings of the data and, in some cases, dramatically different conclusions

as to what the same data mean. For instance, repeated sampling of a single Asian

source population could produce the same distribution of Native American

haplogroups as one larger sampling (Hey 2005). And, as Schroeder et al. (2007)

noted, even the remarkably homogeneous frequency of the 9RA allele in native

North and South American populations does not preclude the possibility of minor

genetic contributions from other groups one or more times in prehistory. I conclude

that based on genetic data alone, the number of founding populations is not

resolvable, or in any event has not yet been resolved. Introducing archaeological

data, however, brings the picture into clearer focus.

In addition to lacking clarity on the numbers of founding groups, molecular

scientists struggle to estimate the number of people who constituted the (or the

various) founding New World populations. Hey (2005) argued that the founding

population was small, on the order of just 70 or so people. Kitchen et al. (2008),

however, who also subscribe to a single-founder peopling model, proposed a

founding population of 640 women, sampled from a source population of

4,000–5,000 women in Siberia/Beringia. Assuming approximate gender parity,

Kitchen et al. (2008) put the total New World founding population at around 1,280

336 J Archaeol Res (2011) 19:327–375

123

Page 11: fulltext (2)

individuals and the northeast Asian source population at 8,000–10,000. Most other

researchers decline to specify an absolute number of founders, leaving us to

consider two scenarios of founding-population size that differ by two orders of

magnitude.

A debate between Anderson and Gillam (2000, 2001) and Moore and Moseley

(2001) focused on the peopling process within the Americas immediately upon

founding offers fodder for evaluating the competing hypotheses of Hey (2005) and

Kitchen et al. (2008). Anderson and Gillam (2000) ran four least-cost simulations,

starting in four possible points of origin in North America, with four destinations, to

evaluate the likeliest route(s) for human colonization of the New World. In various

simulations, they invoked founding population sizes of 25, 50, and 175 people,

respectively, because these figures are widely accepted to represent the population

range of ethnographically and archaeologically known forager bands (e.g., Kelly

1995)—although we are well advised to remember that the colonization of the

Americas involved an unpopulated landscape, which no ethnographic example of

foragers does. While they favored the highest of these figures based on the

pioneering Paleolithic population-size modeling of Wobst (1974, 1976), all of their

demographic parameters were more in line with Hey’s (2005) hypothesized

founding population size of 70 than with Kitchen et al.’s (2008) 1,280 people.

In a pointed response, Moore and Moseley (2001) argued that Anderson and

Gillam had invoked founding population sizes that if not nonviable in the absence of

groups with which to interact and intermarry, were at least highly unlikely to have

been large enough to populate two continents. Moore and Moseley (2001) ran their

own simulations, concluding that all of Anderson and Gillam’s (2000) hypothetical

bands became extinct within 1,000 years under nearly all circumstances modeled.

Anderson and Gillam (2001) acknowledged Moore and Moseley’s (2001) point that

small founding populations are highly likely to have gone extinct, and that some

very early American immigrant groups may have done so. Anderson and Gillam

(2001) pointed out, however, that they had envisioned two scenarios that could have

solved the low-founding population group size problem: the presence of ‘‘macro-

bands’’ of 500–1,500 people (conceptually similar to Wobst’s [1974] ‘‘maximum

band’’ of 175–475 people) or the evolution of ‘‘staging areas’’ where fissioning

bands would have maintained contact with other groups and dramatically increased

the prospects for founding success.

Collectively and despite the apparent contradictions, the work of Hey (2005),

Kitchen et al. (2008), Anderson and Gillam (2000, 2001) and Moore and Moseley

(2001) and the empirical studies that support their assumptions converge to support

a credible peopling scenario. We know that the people who founded and colonized

the New World were highly mobile foragers (e.g., Jones et al. 2003; Surovell 2000).

If they and their immediate descendants had not been, they would not have

populated the New World. We also know that highly mobile forager band sizes

throughout time and across space rarely exceeded the maximum 175 people

modeled by Anderson and Gillam (2000) and were typically smaller (although again

we must recall that an empty continent puts us in uncharted territory). Nonetheless,

we know that maintenance of small population sizes invariably requires mobile

forager bands to interact and exchange mates with others, something the very First

J Archaeol Res (2011) 19:327–375 337

123

Page 12: fulltext (2)

Americans could not have done within the confines of the unpopulated Americas.

Parsimony therefore lends weight both to Moore and Moseley’s (2001) simulation-

based conclusion that a founding population size in the 25–175 range would have

been nonviable and to Anderson and Gillam’s conception of macrobands or a

staging area that would have put the real founding population size in the range of

500–1,500 people.

Given the above, we are left with one scenario that squares with what

anthropologists know about forager behavioral ecology. If a single band of 1,280

people per Kitchen et al. (2008) migrated en masse from Asia to the Americas to

populate the New World, that behavior would have been, at best, incredibly unusual

within the context of ice-age hunter-gatherer demography and mobility. Leaving

aside possible impetuses for such a mass exodus of foragers, the fact that no single

founder could have known that an entire New World awaited colonization, and the

logistics required to amass so many people at one moment in time, we must consider

a scenario more in line with what foragers have been documented to do (see

Surovell [2000] for a relevant discussion). They do travel in band sizes along the

lines of Hey’s (2005) 70 and Anderson and Gillam’s (2000) 25–175 people. And

these are viable population sizes given contact with other groups constituting total

densities in the 1,280 range of Kitchen et al. (2008) and Anderson and Gillam’s

(2000) 500–1,500-person macroband or staging population.

If the necessary population densities for a successful colonization were not

waiting in the New World, which they were not, then they had to have been camped

back home in Siberia, either in western Beringia, which again, some geneticists

have argued hosted a bottlenecked population during and after the LGM, and/or in

Beringia’s original source population in south-central Siberia, fully 4,000 km to the

southwest. A Beringian population not only would have provided a geographically

accessible source of human interaction for an otherwise nonviable founding New

World band but would have produced, if genetically sampled even multiple times,

the array of haplogroups documented today among native North, Central, and South

Americans. The latter, of course, also would have been true of the south-central

Siberian population that spawned both Beringian populations and ultimately New

World colonization. In short, and based strictly on widely accepted parameters of

forager behavior, Hey (2005) is probably right that ‘‘the’’ founding North American

band was the size of a typical forager band. However, it is also highly likely that

Hey’s (2005) founding band maintained gene flow with its source population(s) in

Siberia or with other small bands that migrated from the same source region over a

short time span.

Conceived a different way, and one that embraces the conclusions of so many

geneticists that just one or two populations founded the New World, we must simply

be clear that the ‘‘founding population’’ of the Americas was not, at least not

exclusively, a unit precisely equivalent to the first discrete band of foragers to

immigrate to the new continent. Rather, the founding population came from a larger

source population, along the lines of the ‘‘macroband’’ of Anderson and Gillam

(2000) that for a time at least experienced intercontinental gene flow. If sufficiently

isolated, but not necessarily entirely isolated, from other Asian populations, as

genetic evidence suggests latest Pleistocene western Beringians and, for that matter,

338 J Archaeol Res (2011) 19:327–375

123

Page 13: fulltext (2)

south-central Siberians, probably were (e.g., Starikovskaya et al. 2005), then such a

unit can be reasonably viewed as a single founding population or the source of two

major migrations that could have taken rather different forms.

The view from osteological and archaeological databases

Recent genetic investigations have provided a fertile array of evidence for assessing

Old World geographic origins of First Americans, the time frame for their initial

arrival in the Americas, and the likely structure and size of founding populations

and their migrations. Genetic data are not well equipped, however, to speak to some

crucial issues, including most notably how, specifically, early immigrants made the

voyage from the Old World to New World and how, even generally (e.g., by coast

or via an interior route), they populated the Americas upon arrival. Here I focus on

osteological and archaeological evidence, both in terms of the support it lends or

does not lend to genetic-based inferences and what it suggests about related issues

not informed by extant genetic studies.

Osteological data

The osteological database of First Americans consists of only a very few latest

Pleistocene specimens and a few others dating to the Early Holocene. These finds

have been made over a vast geographic expanse, with the larger percentage and

most ancient specimens representing Central (Gonzalez-Jose et al. 2005) and South

America (Neves and Hubbe 2005; Neves et al. 2003, 2004, 2007; Sardi et al. 2005),

not North America (Jantz and Owsley 2001; Neves and Blum 2000; Owsley and

Jantz 2001). The studies, which focus on craniofacial characteristics, reveal that the

oldest New World human remains show a generalized morphology that aligns them

not with contemporary northeast Asians but with contemporary Africans and

Australians, in the case of South American finds (e.g., Neves and Hubbe 2005;

Neves et al. 2003, 2004, 2007), and east Asians, north Asians, Polynesians, and

Europeans for specimens from North American contexts (e.g., Brace et al. 2001;

Chatters 2000; Jantz and Owsley 2001; Steele and Powell 2002). Researchers agree

that there is a general lack of similarity between the oldest New World remains and

contemporary Native Americans (e.g., Chatters 2000; Powell 2005; Swedlund and

Anderson 2003).

These observations might appear to contradict genetic data showing a straight-

forward relationship between ancient and modern Native Americans and northeast

Asians. Osteologists point out, however, that the generalized appearance of the First

Americans is not unique to that region, but also characterizes early finds from China

(Cunningham and Jantz 2003; Neves and Blum 2000; Neves et al. 2007) and Europe

as long ago as 40,000-24,000 cal. B.P. (Holliday 1999; Jantz and Owsley 2001).

Gonzalez-Jose et al. (2005) interpret this to indicate that the initial colonization of

the Americas preceded the origin of the more specialized skeletal morphology of

northeast Asians, which several researchers have argued emerged around the

Pleistocene/Holocene transition (Cunningham and Jantz 2003; Neves et al. 2003,

J Archaeol Res (2011) 19:327–375 339

123

Page 14: fulltext (2)

2004, 2007; but see a relevant dissertation by Auerbach [2008]). This interpretation

has obvious implications for the timing of the initial peopling of the Americas in

that a migration by people prior to the Pleistocene-Holocene boundary necessarily

pushes the event back to sometime before 13 k cal. B.P. Brace et al. (2001,

p. 10,017), in fact, do propose a peopling date of ‘‘maybe 15,000 years ago,’’ a time

frame consistent with the scenario advocated in this paper.

Some, including Jantz and Owsley (2001), Owsley and Jantz (2001), and Sardi

et al. (2005), interpret the high level of craniofacial variability of the earliest

Americans as evidence for multiple migrations by entirely different populations, a

conclusion that conflicts with the genetic evidence. Neves and Blum (2000) offered

a variation on this interpretation that does not so overtly contradict genetic

observations, suggesting that the craniofacial database of the Americas represents

two closely spaced migratory pulses from northeast Asia. The two migrations, they

argued, could have sampled the same source population, which would have retained

a consistent array of haplogroups through time and contributed that array to

members of both immigrant populations. Neves and Blum (2000) conceived of the

initial colonists as characterized by the highly generalized skeletal morphology seen

among the oldest human remains from South America (Neves and Hubbe 2005;

Neves et al. 2007), followed by immigrants with specialized Mongoloid traits of

contemporary northeast Asians, such as those of 1,000 year-younger remains found

near Buhl, Idaho, that show the latter characteristics.

Archaeological data

Archaeologically, the key to interpreting processes involved in the peopling of the

Americas is fundamentally rooted in the locations and ages of sites in the source

region in the Old World and, of course, the New World. The literature is replete

with manuscripts reporting the excavations and interpretations of early, peopling-

relevant sites on both sides of the Bering Strait, each followed in quick succession

by additional papers critiquing methods, findings, and interpretations. Accordingly,

although I make reference to particular sites here and relate them to my scenario of

the peopling of the New World, I do so knowing that I will mention sites that some

readers reject, fail to include sites that some readers embrace, and ultimately fail to

please anyone. I will attempt, however, to be as transparent as possible about the

degree of acceptance each site enjoys in the archaeological community.

Asia and Beringia

I argue that the First Americans traced their origins to south-central Siberia;

specifically the Transbaikal and Altai Mountain/Yenesei Valley regions, and,

subsequently, western Beringia. Genetic evidence supports this view and also points

to an 18 k–15 k cal. B.P. time frame for the initial colonization of the New World,

together with a population bottleneck in western Beringia at some point during or

just after the LGM. How well does the archaeological evidence mesh with the

genetic picture? In broad terms, rather well, but there is a gap in the data that must

be acknowledged that could contradict an element of the model I advocate, and that

340 J Archaeol Res (2011) 19:327–375

123

Page 15: fulltext (2)

should certainly spur future research. The gap is in the archaeological record of

western Beringia, c. 27,000–14,200 cal. B.P., a period encompassing the LGM and

the 6,000–7,000 years following it. If the lack of western Beringian sites dating to

this time span represents a real lack of human occupation, as some have argued and

I later discuss, this will compel reevaluation of the idea that humans spent a

significant period of time there after leaving south-central Siberia and before

entering the Americas.

Archaeologists agree that humans had settled Siberia, including the far north and

western Beringia, by 32 k cal. B.P. and thus before the LGM (e.g., Goebel et al.

2008; Hoffecker and Elias 2003; Kuzmin and Keates 2005; Pavlov et al. 2001;

Pitulko et al. 2004). Most pertinent to the peopling story is the Yana Rhino Horn site

(Yana RHS), located on the Yana River and south of the Arctic Ocean at 71� N

latitude (Fig. 1). This location puts the Yana RHS within the confines of western

Beringia, defined by Hoffecker and Elias (2003) as the region bounded by the Lena

River or Verkhoyansk Mountains to the west and the Kamchatka Peninsula as far

south as 50� N latitude during the LGM. The single-component Yana RHS produced

Clovis-like carved foreshafts made of wooly rhino horn and evidence for bifacial

chipped stone technologies. The site has been securely dated to 32 k cal. B.P., with

one of many consistent radiocarbon dates derived from a Pleistocene horse bone

with a human-produced flake embedded in it.

Archaeologists also agree that by 14,200 cal. B.P. people had reached the Swan

Point site in central Alaska, eastern Beringia (e.g., Bever 2006; Largent 2004)

(Fig. 2). Swan Point, located in the Tanana Basin, yielded microblades among other

stone implements, together with radiocarbon dates older than Clovis. The nearby

Tanana Basin sites of Broken Mammoth and Mead also have produced radiocarbon

dates in the 14 k–13 k cal. B.P. range, together with blades, burins, scrapers,

bifacial thinning flakes, and mammoth ivory rods (Bever 2001, 2006; Dumond

2001; Hoffecker 2001; Hoffecker and Elias 2003; Holmes 2001; Yesner 2001).

Human occupation dating to the latter portion of this time span has likewise been

well documented in western Beringia, at Ushki Lake on the Kamchatka Peninsula in

the maritime region of southwestern Beringia at 55� N latitude (Goebel et al. 2003)

(Fig. 1). Recent excavations of Ushki Lake’s oldest component 7 produced two

hearths dated to c. 13 k cal. B.P., accompanied by burnt bone fragments and lithic

artifacts on a living surface (Goebel et al. 2003).

Consensus exists that humans occupied western Beringia by 32 k cal. B.P. and

were in the region c. 13 k cal. B.P., and that human occupation of eastern Beringia

began as early as 14.2 k cal. B.P. This leaves the aforementioned gap in human

occupation of Beringia from 32,000 to 14,200 B.P. Some, including Dolukhanov

et al. (2002), Goebel (1999, 2002), Hoffecker (2005), Hoffecker and Elias (2003),

and Mandryk et al. (2001), have argued the gap reflects real depopulation of

northern and central Siberia. Goebel (1999), among others, attributed the inferred

demographic shift to the onset of bitter LGM conditions that forced people to retreat

south to refugia in the Transbaikal and southern Yenesei areas. Kuzmin and Keates

(2005), however, dispute that LGM depopulation of Siberia occurred at all, arguing

that evaluating the number of human occupations, rather than the number of

radiocarbon dates per se, reveals no significant population decrease during the LGM

J Archaeol Res (2011) 19:327–375 341

123

Page 16: fulltext (2)

Fig. 2 Map depicting the Americas and archaeological sites discussed in the text. Circles and numbersrepresent non-Clovis sites (although note that Cactus Hill and Topper have overlying Cloviscomponents). Squares and letters represent Clovis sites. Clovis sites depicted include localities thatFaught (2008), Hamilton and Buchanan (2007), Waters and Stafford (2007a), and I agree definitively dateto the Clovis era (13,250–12,800 cal. B.P.) and have produced in situ Clovis artifacts. (1) Swan Point,Mead, Broken Mammoth; (2) Bluefish Caves; (3) On Your Knees Cave; (4) E1Ta-18 and Namu; (5)Kennewick; (6) Cooper’s Ferry; (7) Indian Sands; (8) Paisley 5 Mile Point Caves; (9) Cross Creek; (10)Arlington Springs, Daisy Cave, Cave of Chimneys, other Channel Island sites; (11) Quebrada Jaguay;(12) Quebrada Tacahuay, Ring; (13) Quebrada Santa Julia; (14) Monte Verde; (15) Clovis-era Patagoniasites; (16) Page-Ladson; (17) Topper; (18) Cactus Hill; (19) Meadowcroft; (20) Hebior, Schaefer; (21)Lovewell; (22) La Sena; (A) Richey-Roberts; (B) Anzick; (C) Colby; (D) Lange-Ferguson; (E) Casper;(F) Dent; (G) Jake Bluff; (H) Domebo; (I) Lehner, Murray Springs; (J) Hiscock; (K) Shawnee-Minisink.Dashed lines depict the approximate extent of the Cordilleran and Laurentide ice sheets c. 14 k–13 k cal.B.P. and the ice-free corridor between them. Figure drafted by Holly Andrew

342 J Archaeol Res (2011) 19:327–375

123

Page 17: fulltext (2)

(see Straus et al. [2000] for a similar argument made in the context of Iberia).

Moreover, Kuzmin and Keates (2005) reported exponential population growth

throughout the region during the warming period that immediately followed the

LGM and extended to the Pleistocene–Holocene transition.

Environmentally, much of Beringia remained ice-free for the duration of the

LGM and hosted expanses of tundra that supported over two dozen megafaunal

species. These included the mammoth, horse, and bison known to have been

economically vital to both Old and New World Upper Paleolithic hunters living

under cold climatic conditions (Barnes et al. 2007; Hoffecker and Elias 2003;

Shapiro et al. 2004). Based on studies of plant macrofossils from, among other

sources, the stomach contents of a frozen extinct horse, Zazula et al. (2003)

characterized the LGM environment of eastern Beringia as largely mammoth steppe

and the vegetation thereof as highly productive dry grassland. In Beringian

lowlands, areas subsequently submerged by rising sea levels of the waning

Pleistocene, wet-adapted plant species were common in the LGM (Hoffecker and

Elias 2003). Brubaker et al. (2005) have similarly documented that parts of Beringia

served as LGM refugia for arctic herbs, shrubs, and boreal trees, species that quickly

recolonized much of the region during terminal Pleistocene warming.

The ecological structure of LGM Beringia did not constitute an obvious

impediment to human occupation during or just after the LGM, particularly given

that Kuzmin and Keates (2005) have demonstrated that similar settings supported

human populations in central and southern and perhaps northern Siberia at that time

(also see Willerselv et al. [2003] for pertinent Siberian paleoenvironmental data

derived from megafaunal DNA preserved in permafrost). Hoffecker and Elias

(2003) offered one possible reason why people may have avoided the Far North and

Beringia during the LGM, if in fact they did: humans then occupying the region may

have retained their warm-climate skeletal morphology, a vestige of their southern

roots that would have made survival in the coldest regions of the world difficult.

However, Siberia has yielded just one site, Mal’ta, with LGM human bone (Kuzmin

and Keates 2005; Kuzmin et al. 2009; Richards et al. 2001), so evaluating this

hypothesis poses a challenge. Even if true, by 21 k-19 k cal. B.P., skeletal finds in

other extremely cold Old World settings show evidence for incipient physiological

adaptations to cold that would have facilitated human expansion by the period

immediately following peak LGM conditions (Hoffecker 2002; Holliday 1999).

The question and classic dilemma thus remains: was Beringia really devoid of a

human presence for the nearly 18,000 years between 32,000 and 14,200 years ago,

or have Beringian sites dating to this period not yet been located? Several

controversial finds hint that such occupation may have occurred. Most convincing is

Bluefish Caves, located along the Porcupine River in the northern Yukon at 67� N

latitude in eastern Beringia (Fig. 2). Morlan (2003) argued on the basis of what he

interpreted to be human-altered mammoth bone that the site was occupied nearly as

early as western Beringia’s Yana RHS and again c. 15,800 cal. B.P. The latter date

has been associated, although neither directly nor unequivocally (e.g., Bever 2001;

Dixon 1999), with an assemblage of wedge-shaped microblade cores, microblades,

burins, and flakes. Morlan (2003) also importantly reported the directly associated

J Archaeol Res (2011) 19:327–375 343

123

Page 18: fulltext (2)

finds of a burin and the remains of species of lemming and vole that lived in the

vicinity of the site only prior to c. 16 k cal. B.P.

In western Beringia, the site of Berelekh (Vereschagin and Ukraintseva 1985;

Fig. 1), located near the mouth of the Indigirka River on the Arctic coast at 71� N

latitude, has long been considered possible—even likely (Pitulko et al. 2004)—

evidence of early, though post-LGM human use of Beringia. Berelekh yielded

mammoth remains reportedly associated with artifacts, including blades, biface

fragments, and stone beads. Wood at the site has been dated to 16,100–15,600 cal.

B.P., although Goebel et al. (2003) and Hoffecker et al. (1993) have questioned the

geologic context of the wood vis-a-vis the mammoth bone and artifacts and

suggested that the mammoth bone could have been scavenged at a later time.

More conservative estimates date occupation of Berelekh to14 k–13 k cal. B.P.

(Hoffecker and Elias 2003), which would make it about the same age as or a little

older than Ushki Lake.

Two other Siberian sites bear mention, Ikhine 2 (Abromova 1989) and Verkhine-

Troitskaya (Mochanov and Fedoseeva 1996), both located in Yakutia along the

Aldan River at 63� N latitude (Fig. 1). Their location puts them at or very near the

westernmost boundary of Beringia at the height of the LGM, and their radiocarbon

dates—24,300–23,400 cal. B.P. for Ikhine 2 and 22,000–21,500 cal. B.P. for

Verkhine-Troitskaya—reflect LGM occupation. However, Kuzmin and Keates

(2005) urge caution when evaluating the dates for both localities. In the case of

Ikhine 2, which yielded assays on wood and bone, they note that the wood could

have been redeposited from younger sediments, while the bone could have been

scavenged later in prehistory. Verkhne-Troitskaya’s single radiocarbon date derived

from wood preserved in permafrost could again have been reworked from deposits

older than the cultural material.

There is, in short, some evidence for human use of late Wisconsinan Beringia

beyond the unequivocal Yana RHS and the Tanana Valley localities. Site dates, if

accepted, narrow if not eliminate the gap, pushing post-LGM occupation of the

region back to about 16 k cal. B.P. or, if one accepts Ikhine 2 and Verkhine-

Troitskaya, to c. 24 k cal. B.P. In addition, Hoffecker and Elias (2003) note that the

lack of modern settlement and development in lands once part of Beringia could

account for the paucity of sites older than 14 k cal. B.P. They may understate the

case. The 2002 Russian census showed that just 6.7 million people occupy the Far

Eastern Federal District (FEFD) that encompasses Beringia and eastern Russia south

to Vladivostok—about 6.2 million km2 of territory. Seventy-five percent of modern

FEFD occupation occurs in nine cities, all but two of which are well south of

Beringia. The exceptions are in southernmost Beringia at its LGM extent: Yakutsk

along the Lena River (population 211,000) and Petropavlovsk-Kamchatskiy on the

Kamchatka Peninsula that hosts Ushki Lake (198,000). The overall population

density of the FEFD today is 1.1 persons/km2, among the lowest densities in the

world. We should probably marvel that we have as robust a Beringian archaeo-

logical picture as we do given how few contemporary Beringians (and Beringian

archaeologists) use that landscape.

Too, contemporary Beringia does not begin to approach LGM Beringia, when the

low sea level exposed vast expanses of land that are now submerged (Fig. 1). None

344 J Archaeol Res (2011) 19:327–375

123

Page 19: fulltext (2)

of us embraces arguments made on the basis of negative evidence; they are

inherently unsatisfying. If we assume for the sake of argument, however, that LGM

sites were sparsely and randomly distributed across Beringia, then 50% or so of

them are, in fact, now under water, regardless of whether that unfortunate reality

makes us uncomfortable. If the sites were disproportionately located adjacent to the

Laptev, East Siberian, Chukchi, Okhotsk, or Bering Seas—which the just-inland

locations of the few documented late Pleistocene western Beringian sites suggests is

actually quite likely (see, for example, the locations of the Yana RHS and Berelekh

in Fig. 1)—they could all be underwater. That evidence of LGM occupation of

Beringia may now be inconveniently located does not logically compel rejection of

the possibility that it exists.

Where does this leave us in terms of where, precisely, the soon-to-be New World

founders spent their time just prior to embarking for the Americas? We have two

choices, and the genetic evidence does not contradict either one, nor can it

contradict either one, because both contenders are clearly part of the peopling

picture. The point of contention is which region hosted the founding New World

source population in sustained fashion immediately prior to colonization. I advocate

Option One, that people never left Beringia and survived LGM conditions in the

region. In so doing, their small numbers and relative isolation created the genetic

bottleneck molecular data suggest existed. That populace stood geographically

poised to enter the New World just as soon as post-LGM recession of the

Cordilleran and/or Laurentide Ice Sheets permitted movement across Beringia and

south into the lower 48 states.

Option Two, favored by Goebel et al. (2008) among others, views people as

having retreated from northern Siberia and western Beringia during the worst of the

LGM and returned to refugia in the Transbaikal and Yenesei Valley portions of

south-central Siberia. There they remained, 4,000 km southwest of the gateway to

the Americas, until improved Beringian conditions beckoned an enterprising

population to penetrate farther east in Beringia than ever before and to continue

south in the New World. In fact, Kuzmin and Keates’ (2005) work references a half-

dozen well-dated LGM sites in south-central Siberia, including Mal’ta, with

radiocarbon-dated human remains (Richards et al. 2001) (Fig. 1). The sites leave

little doubt that south-central Siberia saw LGM occupation. The area today falls

within the Siberian Federal District (SFD), a 5.1-km2 territory that according to the

2002 census supports over 20 million people, mostly in cities clustered in the

vicinity of the six well-dated LGM sites (e.g., Irkutsk, population 583,000, and

Ulan-Ude, 353,000). The population density of the SFD today is 3.9/km2, four times

that of the FEFD that comprises Beringia. It is plausible, even likely, that the greater

incidence of LGM sites in south-central Siberia vis-a-vis Beringia is a function of

the degree of contemporary development and population density in each, i.e.,

recovery bias.

The Americas

Whether from south-central Siberia without a significant stop in western Beringia or

from a long-occupied western Beringian staging area, humans entered the Americas

J Archaeol Res (2011) 19:327–375 345

123

Page 20: fulltext (2)

via the Bering Land Bridge or the seas adjacent to it. This brings us to the migration

route(s). Paleoenvironmental, glacial, and other studies indicate that the first fully

ice-free corridor followed the coast along the Pacific Rim south along outer Alaskan

shores to the continental United States, and that it was open and populated by plants

and animals by 16 k-15 k cal. B.P. (e.g., Clague and James 2002; Dixon 2001;

Dyke 2004; Fedje and Josenhans 2000; Fedje and Southon 2003; Fedje et al. 2004;

Hoffecker and Elias 2003; Jobling et al. 2004; Leonard et al. 2000; Mandryk et al.

2001; Ramsey et al. 2004; Shapiro et al. 2004). The interior ice-free corridor, which

formed when the Cordilleran and Laurentide Ice Sheets receded (Fig. 2), did not

open and attain ecological viability such that people could have traversed it until

14,000–13,500 cal. B.P. (e.g., Arnold 2002; Burns 1996; Hoffecker and Elias 2003;

Mandryk et al. 2001; Wilson 1996; maps presented by Dyke [2004] show the

possibility of an open corridor as early as 15,200 cal. B.P., but viability for human

passage at that early time is unlikely). Would-be founders of the New World, then,

could have accessed the Americas via an open coastline at least 1,000 years before

the IFC offered an alternative route south. But did they?

Since Fladmark (1979) published the first robustly conceived hypothesis for a

Pacific coast peopling, the most significant impediment to its acceptance has been

the lack of very early sites in coastal settings anywhere along the Beringian,

Alaskan, or Pacific Northwest coastlines, coupled with hard-to-shake misconcep-

tions about the quality and desirability of coastal resources (see Erlandson et al.

[2007] for examples of the latter). Researchers universally recognize the challenge

of demonstrating the presence of Late Pleistocene sites along the coast: if sites were

present, they are now under water (e.g., Clark and Mix 2002), the same dilemma

faced by now-hypothetical lowland Beringian sites of the LGM. Some see this

argument by negative evidence as a problem so profound that we must exorcise the

very notion from our list of peopling scenarios (e.g., Marshall 2001; Turner 2003;

Yesner 2001; Yesner et al. 2004). Others view the idea as worthy of consideration

alongside other peopling models and are developing innovative ways to evaluate it

(e.g., deFrance et al. 2001; Dixon 1999, 2001; Erlandson et al. 2005; Fitzgerald and

Jones 2003; Goebel et al. 2008; Jones et al. 2002; Kelly 2003; Montenegro et al.

2006; Rick et al. 2001; Surovell 2003; Vallanoweth et al. 2003).

For many years, the strongest archaeological evidence for a very early coastal

peopling came from the Monte Verde site, located along Chinchihuapi Creek in

south-central Chile, 58 km inland from the Pacific Ocean (Dillehay 1999) (Fig. 2).

Radiocarbon dating of a variety of materials found in clear association with human-

produced artifacts revealed occupation c. 14,600 cal. B.P.—400 years earlier than

Swan Point in eastern Beringia and a millennium before Clovis. The site convinced

most archaeologists (Meltzer et al. 1997; cf. Fiedel 2000) that people had reached

the Americas before Clovis time, and many viewed its geographic location as

consistent with migration along the west coast of the Americas by boat. Boats, the

rationale goes, would have offered a much faster mode of transport to this far-

southerly site than traveling by foot, presupposing a Beringian origin in either case.

New evidence from the site (Dillehay et al. 2008)—the recovery of nine species of

marine algae and seaweed representing different seasons and coastal settings—

suggests Monte Verdeans exploited marine resources year-round. While this does

346 J Archaeol Res (2011) 19:327–375

123

Page 21: fulltext (2)

not definitively indicate that site occupants reached Monte Verde by watercraft, it

does demonstrate intimate knowledge of the sea and its resources.

In the past decade, archaeologists have published new data on others sites located

along the Pacific Coast that like Monte Verde are more consistent with a coastal

peopling model than one invoking the IFC as the initial conduit for a southerly

pedestrian migration of First Americans (Fig. 2). The sites range in age from Clovis

contemporaries to Early Holocene locales, with nearly all showing subsistence

strategies that revolve around a coastal resource base and not the terrestrial one

invariably attributed to Clovis people (e.g., Byers and Ugan 2005; Cannon and

Meltzer 2004; Kooyman et al. 2006; Redmond and Tankersley 2005; Surovell and

Waguespack 2009; Waters et al. 2009a). Of these sites, Arlington Springs is most

significant, because the site yielded human remains recently dated to

13,100–13,000 cal. B.P., a time frame coincident with Clovis, and because it is

located on Santa Rosa Island, 70 km south-southwest of Santa Barbara, CA

(Agenbroad et al. 2005; Johnson et al. 2002, 2007). Arlington Springs could have

been accessed only by boat, and the site therefore offers the most direct evidence

available that First Americans as early as Clovis time employed watercraft.

South American sites enjoying consensus acceptance and contributing to a

scenario of an early Pacific Coast peopling include Quebrada Jaguay (Sandweiss

et al. 1998), Quebrada Tacahuay, and the Ring site on the southern coast of Peru.

Initial use of Quebrada Jaguay, which dates to 13 k cal. B.P., again coeval with

Clovis, focused on the harvest of ocean fish. Quebrada Tacahuay, 300 years

younger, also reflects procurement of marine resources, in this case cormorants and

boobies, marine fish and mammals, and shellfish (deFrance et al. 2001; Keefer et al.

1998). The Ring site, 20 km north of Quebrada Tacahuay and similarly dated,

shows intensive exploitation of shellfish (Sandweiss et al. 1989). Based on these

sites, deFrance et al. (2001; also see de France et al. 2009) argued that Pleistocene

Peruvians possessed sophisticated knowledge of sea resources, as a coastal peopling

model would predict. The recently published site of Quebrada Santa Julia (Jackson

et al. 2007), located north of Monte Verde in Chile and dating to 13,100 cal. B.P., is

another example of a terminal Pleistocene South American site. Despite being

located on the coast, Quebrada Santa Julia residents procured not sea resources but

now-extinct horses.

Less well known, at least to North American-focused scholars, but significant in

their number, setting, and generalized subsistence focus are a half-dozen Clovis-

contemporary sites in Patagonia, situated along major rivers draining to the Atlantic—

not Pacific—Ocean (Borrero 1999; Lavalle and Bahn 2000; Mena et al. 2003; Miotti

2003; Miotti and Salemme 2003; Miotti et al. 2002). The geographic distribution of

the sites prompted Miotti (2003; also see Miotti and Salemme 2003 and Scheinsohn

2003) to suggest that human colonization of southernmost South America followed

both South American coasts and proceeded inland along rivers, either simultaneously

or with advancement along the Atlantic Coast preceding that along the Pacific. Miotti

(2003) noted, too, that evidence of Pleistocene occupation along the Atlantic Coast

itself, if it exists, would likely now be submerged by the Argentinean Sea.

With the exception of Clovis-contemporary Arlington Springs, sites located

along the Pacific Coast of North America date to the Early Holocene. Rick et al.

J Archaeol Res (2011) 19:327–375 347

123

Page 22: fulltext (2)

(2001) reported that Daisy Cave, also in the Channel Islands, yielded evidence for

hook-and-line fishing and intensive fish exploitation by 11,500 cal. B.P. They

suggested (Rick et al. 2001), too, that the marine environment of the Channel

Islands, with its kelp beds, rocky reefs, and estuaries, is one of the most productive

ecosystems on earth, and that people thrived on its myriad resources from the

moment they encountered them (see Erlandson et al. [2007] for more on this point).

That moment, they further argued, must have predated the well-documented 12 k–

10 k cal. B.P. occupations of Daisy Cave and nearby sites, given the sophisticated

technologies—seaworthy boats, nets, fishhooks, and so forth—already in use by

then. Similarly, Erlandson et al. (2005) documented occupation of the southern

coast of San Miguel Island, 40 km from the mainland and the westernmost of the

Channel Islands, by 9,600 cal. B.P. And Vallanoweth et al. (2003) documented

occupation of the same island c. 8,200 cal. B.P. at Cave of the Chimneys on its

northeast shore. In both cases, San Miguel residents exploited mussels, abalone and

other shellfish, fish, sea mammals, and sea birds; as in the cases of Arlington Springs

and Daisy Cave, they could have accessed the sites only by boat (see Jodry [2006]

for a thorough and thoughtful contemplation of the likely use of watercraft by First

Americans and Fagan [2004] for a discussion of the likely nature of early North

American vessels).

North of the Channel Island sites on the central California coast, Jones et al.

(2002) documented a large shell midden and ground and chipped stone tools at the

Cross Creek site, radiocarbon dated to 10,350–9,700 cal. B.P. They concluded that

people with relatively low mobility occupied the site and did so most intensively in

the spring. They also argued (Jones et al. 2002, p. 215), that ‘‘early findings from

Cross Creek, Daisy Cave, and Arlington Woman reflect an adaptive outgrowth from

a separate coastal migration corridor employed by people with (a) maritime

gathering adaptation.’’ This conclusion earned them a rebuke from Turner (2003),

who views the site as too recent to illuminate peopling processes and too sparse in

food refuse to speak to an economic focus. Fitzgerald and Jones (2003) responded

that Turner had cited outdated evidence but agreed that documenting earlier coastal

sites would advance the coastal peopling cause. They also, however, reiterated their

original point that sites like Cross Creek serve as legitimate, albeit indirect,

evidence for a coastal migration corridor.

Still farther north, the Indian Sands site on the southern coast of Oregon offers

that state’s earliest evidence for coastal occupation (Davis et al. 2002-2004). A

radiocarbon date on dispersed charcoal from a floor that produced 136 chipped stone

artifacts suggests human use of the site as early as c. 12,500–12,150 cal. B.P. More

intensive use of the site, as evidenced by a substantial presence of burned and

unburned mussel shells, dates to c. 9,500–8,500 cal. B.P., after sea levels had risen

and the site was situated within 0.5–0.25 km of the Pacific shoreline (1.0–1.5 km

closer than it had been during the previous occupation). Finally, and farther north

still along the coast of central British Columbia, Cannon (2000) dated shell midden

site EITa-18 to 11,600–11,250 cal. B.P. The site is one of several Early Holocene

localities in the vicinity—11,600–10,700-year-old Namu is another—that collec-

tively reflect a maritime economy that developed ‘‘early and rapidly’’ on the

Northwest Coast (Carlson 1998, p. 23).

348 J Archaeol Res (2011) 19:327–375

123

Page 23: fulltext (2)

Although the past decade of archaeological field and lab work has not given

advocates of the coastal migration hypothesis a ‘‘magic bullet’’ site in a Beringian or

Northwest Coast setting—one dating to the chronological window between the

opening of the coastal corridor and the Clovis era when people ubiquitously

occupied North and Central America—it yielded something close. Most researchers

willing to entertain the coastal migration hypothesis envision New World founders

traversing the Pacific Rim and heading south along the coast by boat, yes, but also

as having moved inland via productive river corridors that allowed them to pursue a

familiar lifeway. Kelly (2003), for example, argued that if coastal migration

occurred, then by the time people reached Chile’s Monte Verde, they should also

have budded off into interior North America. Fix (2005) simulated a coastal

peopling by examining founding haplogroup distributions, concluding they

represent a primary coastal focus coupled with a weak pulse inland. Surovell

(2003), also simulating coastal peoplings, concluded that if the coast served as the

initial migratory corridor, continental waterways like the Columbia River would

have promoted movement inland. He further predicted that such a peopling would

have left evidence of early sites above and below sea level in the Northwest, near

the point of New World entry.

Terminal Pleistocene sites below sea level in the region have notoriously eluded

archaeologists. However, Gilbert et al. (2008b) published convincing evidence of a

very early site above sea level that is consistent with a model of coastal people

moving inland along major rivers in the Northwest. The evidence comprises human

coprolites dated 14,270–14,000 cal. B.P. recovered from Paisley 5 Mile Point Caves

in south-central Oregon (Fig. 2). Although desert today, 14,000 years ago the site

overlooked pluvial Lake Chewuacan and could have been accessed from the coast

via the Klamath and Sprague River corridors. mtDNA analysis of the coprolites

indicated they represent haplogroups A2 and B2, two of the founding Native

American lineages. Poinar et al. (2009) and Goldberg et al. (2009) challenged

Gilbert et al.’s (2008b) interpretations on several grounds, including the potential

for contamination and methodological considerations; however, Gilbert et al. (2009)

and Rasmussen et al. (2009) convincingly addressed each critique. While further

details such as dietary inferences may be forthcoming, analyses offered to date

demonstrate strong evidence of inland penetration of North America 1,000 years

before Clovis by people with none of the technologies associated with the Clovis

tool kit (Beck and Jones 2010). Other sites, such as possible Clovis-contemporary

Cooper’s Ferry (Davis and Schweger 2004), located on the Salmon River and also

devoid of a Clovis tool kit, may represent the same sort of inland movement a little

later in time.

We have, in 2010, a robust suite of paleoecological data that demonstrates that

the coastline of the northern Pacific Rim and Northwest Coast deglaciated

1,000 years or more before the formation of an inland IFC, and a burgeoning

understanding of just how highly productive coastal resources along the Pacific Rim

and West Coast were as early as 16 k cal. B.P. (Erlandson et al. 2007, 2008). There

also is a significant body of archaeological evidence indicating that founding

population(s) availed themselves of the coastal route as soon as they could navigate

it. The suite of evidence includes data points that can be considered direct indicators

J Archaeol Res (2011) 19:327–375 349

123

Page 24: fulltext (2)

of coastal migration and others that are indirect but convincing. Direct evidence

includes the unequivocal residue of human occupation near the coast at Monte

Verde and along a northern coastal waterway at Paisley Cave V prior to 14 k cal.

B.P. This time frame predates the availability of the IFC and rules out the possibility

that people accessed either locality via that route. Arlington Springs Woman also

serves as direct evidence of a coastal peopling, because she reached her final resting

place in Southern California’s Channel Islands by watercraft—and she did so c.

13,100 B.P., only shortly after the IFC had opened for migratory business.

Indirect evidence for coastal migration derives from up and down the Pacific

Coast of North, Central, and South America, and the Atlantic Coast of South

America, where very early sites cluster and almost always reflect at least a part-time

marine orientation. If Cross Creek were the only such site, we might ascribe some

weight to Turner’s (2003) titular objection that ‘‘three ounces of sea shells and one

fish bone do not a coastal migration make.’’ But Cross Creek is far from the only

such site in the Americas; it is one of dozens (Erlandson et al. 2007). Even the few

early human bones from coastal or coastal river corridor contexts bolster the picture

of an early maritime adaptation in the New World. OYKC Man of Prince Wales

Island, 10,300 years old, whose haplogroup D mtDNA profile represents a founding

lineage, ate a marine diet (Kemp et al. 2007). Kennewick Man (Chatters 2000;

Huckleberry et al. 2003), buried and recovered in a bank of the Columbia River,

consumed 70% salmon c. 9,600–9,300 cal. B.P. Buhl Woman, found along the

Snake River that drains to the Columbia, also ate a diet that included anadromous

fish 12,800–12,400 cal. B.P. (Green et al. 1998). People in the New World

possessed intimate knowledge of the sea and other aquatic environments; they

exploited those resources beginning over 14,000 years ago and continued to do so

ubiquitously thereafter. This reality cannot be obscured by high sea levels, and it is

utterly inconsistent with pedestrian colonization of the New World via the IFC.

That said, one might ask why, if terminal Pleistocene coastal environments were

so rich and given the low residential mobility of ethnographically documented

maritime hunter-gatherers (Kelly 1995; Yesner 1980), we should expect First

American founders to have moved so quickly south along the coast. Why not simply

settle into the Pacific Northwest and enjoy the population growth coastal

environments could and later did support? The answer may be rooted in the

question. The First Americans differed fundamentally from ethnographically known

hunter-gatherer populations, who were always documented at or near the carrying

capacities of the environments they occupied (Steele et al. 1998). The First

Americans were first, and the landscape the encountered was, by definition, devoid

of other people.

Recent demographic modeling of migrations of First Americans (Hazelwood and

Steele 2003, 2004; Lanata et al. 2008; Peros et al. 2010; Steele 2009; Steele and

Politis 2009; Steele et al. 1998) and hunter-gatherers entering other unoccupied

regions of the world (e.g., Field and Lahr 2006; Field et al. 2007) indicates that

human colonization of productive, open landscapes occurs quickly. Ray and

Excoffier (2009) have further shown that long-range migration, as from the Pacific

Northwest to South America, tends to accelerate colonization rates. Meltzer (2003),

in this vein, has remarked that on the unexplored landscape of the Pleistocene New

350 J Archaeol Res (2011) 19:327–375

123

Page 25: fulltext (2)

World, selection likely favored rapid and extensive exploration (also see Beaton

1991), and Surovell (2000) noted that at least two later coastal migrations across the

Arctic coasts of northern Alaska and Canada occurred extremely quickly, although

for reasons that probably differed from those explaining rapid coastal migration by

First Americans. Both of these robustly modeled predictions and archaeological

precedents bolster a scenario of rapid colonization of the west coast of the

Americas, despite some arguments to the contrary (e.g., MacDonald 2004).

While I do believe the evidence supports a coastal colonization of the Americas, I

reject the notion that the early occupants of the Pacific Coast spawned all human

occupation of the Americas generally or Clovis specifically. Hamilton and

Buchanan (2007) analyzed the spatial distribution of Clovis-era radiocarbon dates

across North America and evaluated competing models of New World colonization.

Their best-fit model, and the only statistically significant one, initiated colonization

near Edmonton, Alberta, at the presumed mouth of the IFC, and then saw people

move rapidly south and east through North America. The location of Clovis sites

depicted in their map (Hamilton and Buchanan 2007, p. 15627, Fig. 1), and a

number of others like it (e.g., Kelly 2003, p. 140, Fig. 2), supports the idea that

Clovis ancestors arrived via Beringia and the IFC shortly after the corridor opened.

Buchanan and Collard (2007) derived similar conclusions from a statistical analysis

of fluted points recovered from across North America. Cumulatively, these findings

suggest that in some areas Clovis people whose ancestors entered the New World

via the IFC were, in fact, first (see Waguespack [2007] for a similar perspective).

Other archaeologists share the view that colonization of the Americas involved at

least two major migratory events spaced 1,000 years apart. Mandryk et al. (2001)

posited that the first people to populate the Americas did so via the Pacific Coast,

but that their predominant migration down the coast and only slowly inland left the

interior of the continent open for the arrival of people via the IFC. Those later

arrivals, they argued, developed the Clovis tool kit and dispersed the technology

across the Americas. Erlandson (e.g., 2002), too, expressly noted the likelihood that

both the coast and IFC played crucial roles in the peopling process. Beck and Jones

(2010) recently and convincingly argued that very old stemmed point traditions of

the Intermountain West constitute the archaeological residue of coastal colonists

who moved inland prior to Clovis. Faught (2008) envisioned multiple colonization

events, including but not exclusively from the Pacific Rim and the IFC via Beringia.

Even Fiedel (e.g., 1999, 2002, 2004, 2005), who advocates Beringia and the IFC as

the major conduit for immigration of those most responsible for colonization of the

Americas, does not rule out that some people may have arrived earlier via the coast.

More recently, Waguespack (2007) outlined two models for colonization, one pre-

Clovis and coastal, the other proto-Clovis and inland, although she frames them as

mutually exclusive. Goebel et al. (2008), while favoring a single-founder model of

New World colonization, acknowledge that Clovis could represent a second

Beringian colonization event.

Just as the Late Pleistocene–Early Holocene subsistence trajectory of those who

occupied sites along the Pacific Coast of the Americas indicates a focus on maritime

resources from the outset and through time, so too does the evidence of Clovis

subsistence unequivocally reflect a terrestrial hunting focus. A spirited debate has

J Archaeol Res (2011) 19:327–375 351

123

Page 26: fulltext (2)

explored the question of whether Clovis hunters specialized in mammoth and

mastodon hunting (Dunbar 2006b; Surovell 2000; Surovell and Waguespack 2008,

2009; Waguespack 2005; Waguespack and Surovell 2003) or engaged in a more

generalized—but still terrestrial—hunting and gathering subsistence strategy (Byers

and Ugan 2005; Cannon and Meltzer 2004; Dixon 2001; Elston and Zeanah 2002;

Fiedel 2000, 2005; Meltzer 2004, 2009). However, no archaeologist views Clovis as

remotely maritime oriented, nor is there any evidence for this subsistence pose by

Clovis people whose ubiquitous technological detritus reveals they were fully aware

of the presence the Atlantic and Pacific Oceans and the Gulf of Mexico (see

Anderson’s important and ever-growing ‘‘Paleoindian Database of the Americas,’’

an online catalogue of North American sites older than 10,000 years [Anderson

2005; Anderson and Faught 2000; Anderson et al. 2005, 2010]).

Discussion

By 13,100 cal. B.P. two migratory waves of hunter-gatherers with fundamentally

different lifeways are visible in the archaeological record at the continental scale. This

observation corresponds with the conclusions discussed previously of many

molecular scientists who interpret their genetic data sets as reflecting two distinct

migrations rather than a single founding event. To recall two examples, Perego et al.

(2009) analyzed two rare mtDNA haplogroups in American and Asian populations

and concluded that only two independent migrations—one along the coast and one in

northern North America—can account for their current distribution among contem-

porary Native Americans. Schurr (2004) similarly argued based on mtDNA and NRY

evidence that the initial colonists followed a coastal route and expanded from there

into all continental regions, and that a second migration sampling the same Siberian

source region entered the New World shortly thereafter via an interior route.

Schurr’s (2004) interpretation may hold the key for reconciling the argument of

many other molecular biologists that genetic data, namely, a perceived lack of

genetic diversity and a homogeneous intercontinental haplogroup distribution

among contemporary Native American populations, indicate colonization by a

single founding group. As Hey (2005) and others have noted, it is difficult to discern

from any contemporary First American genetic data whether low diversity and

haplogroup ubiquity definitively indicate the migration of one small founding group

versus multiple closely spaced migrations that originated from the same source

population. The peopling scenario I advocate envisions the LGM source population

for New World founders as having occupied Beringia for thousands of years before

a subset embarked on the final push to the continental United States. I also argue

that general principles of forager mobility argue against a mass exodus from

Beringia (or from 4,000 km southwest, for that matter), that would have somehow

changed the overall genetic signature of the source population and/or depleted it.

Archaeological evidence shows that in fact a population ‘‘explosion’’ occurred in

Siberia starting shortly after the LGM (Kuzmin and Keates 2005, p. 785) and that

Beringia acquired a very ‘‘complicated’’ archaeological record by 14 k cal. B.P.

(Goebel et al. 2008, p. 1498). Mandryk et al. (2001) and Fladmark (1986),

352 J Archaeol Res (2011) 19:327–375

123

Page 27: fulltext (2)

moreover, offered the important reminder that Beringia had coasts and a vast

interior, both of which paleoecological studies show were viable landscapes for

LGM human occupation. The different environments would have promoted

different forager subsistence strategies, however, as coasts and homogeneous

grasslands around the world and throughout time always do and have. That LGM

Beringians, i.e., the New World source population, practiced variable subsistence

strategies depending on where they lived, however, does not also suggest that they

differed from one another genetically. Extant evidence indicates the opposite, that

Beringians traced their immediate ancestry to south-central or southeastern Siberia,

where all the New World founding haplogroups have been documented today.

Moreover, LGM populations of Beringia must have been very low—they remain

among the lowest in the world today even under milder conditions—suggesting that

gene flow among occupants was highly likely if not a biological imperative for the

survival of the collective Beringian populace.

To follow this chain of logic to its conclusion, and if at least most of the premises

I have outlined are accurate, then the second migration that is required to account

for the distribution of Clovis sites in North and Central America should not, for two

reasons, be particularly obvious in the genetic signatures of descendent populations.

First, proto-Clovis immigrants sampled the same Beringian source population (or

south-central/southeastern Siberian for those who prefer to forego LGM Beringian

occupation) as their coastal predecessors. Second, gene flow among First Americans

who arrived via the coast and via the IFC was undoubtedly reinitiated as soon as

people re-encountered one another. This would not have taken long, given that we

know that coastal immigrants followed continental waterways east to Paisley Cave 5

over 14 k cal. B.P., and that terrestrial hunting and gathering-oriented Clovis bands

reached East Wenatchee, located along the Columbia River in Washington, by

13 k cal. B.P. (Haynes et al. 2007; Waters and Stafford 2007a, b).

If it is true that we should not expect to readily detect discrete migration events

from Beringia to the Americas in genetic data sets, why then do some geneticists

perceive precisely that? The answer lies at least in part in the nature of the data

themselves. Perego et al. (2009) focused on two rare Native American mtDNA

haplogroups (D4h3 and X2a) that have much more discrete geographic distributions

today than do the four major founding haplogroups. Theirs is an approach that

O’Rourke (2009) endorses as likely more informative than the more frequent focus

on the most common and widely distributed founding haplogroups. Schurr (2004)

similarly noted that NRY haplogroups P-M45b, C-M130, and R1a1-M17 appear to

have been disseminated only in North and Central America, coincident with the

opening of the IFC and occurring after dissemination of the major mtDNA

haplogroups and NRY P-M45a and Q-242/M3, which dispersed throughout the New

World after a coastal entry.

Problems with the proposed dual-migration model

In the Introduction I suggested that scientists synthesizing data sets to create

peopling narratives point out the gaps and weaknesses in their own scenarios. First,

J Archaeol Res (2011) 19:327–375 353

123

Page 28: fulltext (2)

this might add a dose of humility to the enterprise that could help us move beyond

the gratuitous nastiness that surfaces increasingly often in peopling-related

manuscripts. Adovasio and Page (2003), Fiedel (2002), and Grayson and Meltzer

(2003) are just a few of the otherwise outstanding practitioners whose occasionally

toxic rhetoric undermines their scientific positions and repels some readers. Even

more important, the weaknesses and gaps in any synthesis point unerringly to the

subjects in greatest need of additional consideration and research.

I already identified one problem with the scenario I advocate, that the Beringian

archaeological record does not unequivocally support the argument that soon-to-be

New World founders occupied Beringian plains and coasts during the LGM. Some

genetic data suggest this occurred, and paleoecological reconstructions of the

Beringian landscape indicate it could have occurred. But ‘‘could have occurred’’ is

not good enough, and the genetic data also can be construed to support an LGM

retreat back to south-central Siberia’s Transbaikal and Altai regions. On the other

hand, and as discussed at some length, there are valid reasons to expect that

Beringia will yield such sites and reasonable explanations for why it has not done so

yet. So, intrepid field archaeologists, peopling glory may await you in the still-

exposed regions of LGM Beringia and, for that matter, on its now-submerged LGM

shores.

Other data points that challenge my interpretation of the peopling process are

located on the other side of Bering Strait in the Americas. Any synthetic peopling

undertaking requires citation of archaeological sites that support the case being

made. This invariably leads to charges that the synthesizer has cherry-picked those

sites that support his or her perspective on the big picture and ignored those that do

not. One Paleoindian archaeologist’s ‘‘cherry-picking,’’ after all, is another’s well-

reasoned acceptance of some very early sites and rejection of others. I have tried to

incorporate into my narrative sites that I think are best supported by the evidence

and that enjoy something approaching consensus acceptance (and to be transparent

when consensus has not been achieved). Sites mentioned to this point bolster my

narrative. However, a few potentially, even convincingly, very old sites not yet

mentioned do not, and while they have each been challenged on various grounds,

dismissing them out-of-hand is unjustified.

Sites that I think most powerfully challenge my peopling scenario include

Page-Ladson on the Gulf Coast; Topper, Cactus Hill, and Meadowcroft on the

Eastern Seaboard; Hebior and Schafer in America’s Heartland; and the La Sena

and Lovewell localities of the High Plains (Fig. 2). Of these, only Page-Ladson

can, without truly absurd machinations, be reconciled with the scenario I have

presented, and even then the fit is imperfect. Page-Ladson (Dunbar 2006a, b) dates

to 14,400 cal. B.P. and consists of a human-altered mastodon tusk associated with

a few chipped stone artifacts, all from a buried context now submerged by

Florida’s Aucilla River. Faught (e.g., 2002-2004, 2004, 2006) and Lewis (2000)

have argued convincingly that depending on the mechanism, submersion can

cause less damage to sites than erosion and other invasive terrestrial processes,

and Page-Ladson may thus have been fortuitously preserved.

First Americans who colonized the Americas via the Pacific Coast using boats

could have accessed Page-Ladson by crossing Central America at the Isthmus of

354 J Archaeol Res (2011) 19:327–375

123

Page 29: fulltext (2)

Tehuantepec (or South America at the Isthmus of Panama) and circumnavigating

the Gulf of Mexico to the Aucilla River. The distance required to reach Page-

Ladson from the Gulf of Tehuantepec would have been far less than the distance

people definitively navigated to reach Monte Verde two centuries later. However,

the people of Page-Ladson hunted or processed a mastodon, not marine resources.

For at least that moment in time, First Floridians were not evidencing a maritime

orientation as the Late Pleistocene people of the Pacific Coast did and as we would

expect ancient mariners to do. Faught (2008) resolves the dilemma of Page-

Ladson’s early age and geographic position by permitting an early Atlantic crossing,

a possibility favorably modeled by Montenegro et al. (2006) and advocated by

Bradley and Stanford (2004), but contradicted by extant genetic evidence. Straus

(2000), Straus et al. (2005), and Meltzer (2004) also convincingly rejected the

Atlantic peopling model on archaeological grounds. So Page-Ladson can perhaps be

reconciled with the peopling model I advocate if we accord early coastal residents

the latitude to occasionally hunt terrestrial megafauna, as occurred at Peru’s

Quebrada Santa Julia 1,300 years later. However, the reconciliation must stop

there—and should perhaps exclude Page-Ladson—for reasons related to the

chronology, geography, and lifeways represented by the sites that most compel-

lingly challenge my peopling perspective.

South Carolina’s Topper site and Virginia’s Cactus Hill are both located on the

banks of rivers that drain to the Atlantic Ocean, both yielded absolutely dated Clovis

components, and both have components underlying Clovis dated by radiocarbon or

OSL to c. 20,000-16,000 years ago (Goodyear 2006; Goodyear and Steffy 2003;

Macphail and McAvoy 2008; McAvoy and McAvoy 1997; Steffy and Goodyear

2006). Geologists (Feathers et al. 2006; Wagner and McAvoy 2004; Waters et al.

2009b) have convincingly argued for the integrity of the pre-Clovis sediments at both

sites, and Goodyear (2006) illustrated microblades, blade cores, burins, and other

chipped stone tools from both sites that natural forces would have difficulty

replicating. Meadowcroft Rockshelter paints a similar picture, located in western

Pennsylvania’s Cross Creek drainage, a tributary of the nearby Ohio River. Unlike the

Savannah and Nottoway Rivers of Topper and Cactus Hill, the Ohio drains west to the

Mississippi, but the proximity of all three sites to major river corridors is notable. A

minimum date on carbonized basketry places humans at Meadowcroft by 18,500 cal.

B.P., and dates of 15,200–13,400 cal. B.P. are associated with small prismatic blades,

blade cores, and lanceolate projectile points (Adovasio et al. 1999).

Neither the Schaefer and Hebior nor Lovewell and La Sena sites are located

along a major river corridor. Less than 30 km from the southwestern shore of Lake

Michigan in extreme southeastern Wisconsin and just 2 km apart, Schaefer and

Hebior indicate mammoth hunting or scavenging 14,800–14,200 cal. B.P. (e.g.,

Falk 2009; Joyce 2006; Joyce and Blazina-Joyce 2002; Marshall 2001; Overstreet

and Kolb 2003). That time frame, replicated via radiocarbon on multiple samples of

mammoth bone and charcoal, makes the sites contemporaries of Monte Verde,

Paisley Cave V, Page-Ladson, and Swan Point. Researchers documented human-

produced blades at Schaefer and evidence of human butchering of both mammoths

immediately after death (Overstreet 2005). Lovewell and La Sena also represent

finds of human-altered mammoth bones, these in north-central Kansas and

J Archaeol Res (2011) 19:327–375 355

123

Page 30: fulltext (2)

south-central Nebraska, respectively (Holen 2006). Both mammoths were deeply

buried in loess and exhibited spirally fractured limb bones and evidence for

intentional flaking of fresh mammoth bone. Holen (2006) also recovered a highly

polished bone object, reminiscent of a Clovis or Upper Paleolithic bone rod tip,

among fractured bone fragments at Lovewell. Mammoth bone at both sites

consistently dates from 22 k–19 k cal. B.P., much older than Schaefer and Hebior

and overlapping the latter half of the LGM.

As mentioned, the aforementioned sites have been challenged on various

grounds, some vociferously. Goebel et al. (2008, p. 1500) summarized the

challenges: insufficient reporting for Page-Ladson, potential natural origin of

alleged stone tools in early Topper levels, charcoal translocation at Cactus Hill,

contamination at Meadowcroft, and the possibility that natural forces caused the

purported human alteration of mammoth bone at Hebior, Schaefer, La Sena, and

Lovewell. In each instance, however, the researchers responsible for the excavations

and interpretations have meticulously presented their cases, and I suspect that if they

had made the very same cases on a noncontroversial stage, none of the sites would

have been challenged at all. It is worthwhile, then, to explore scenarios that could

account for these sites, again under the premise that what defies explanation or tidy

peopling models should guide future research.

Importantly, none of aforementioned sites is particularly consistent with a coastal

colonization model, whether via the Pacific Rim as argued here, the Atlantic Ocean

(Bradley and Stanford 2004), or even the South Pacific (Wyatt 2004). The northern

Pacific Coast peopling model rests on clear evidence for chronologically early Pacific

coastal and Pacific waterway sites, demonstrably early use of watercraft, and a very

long-standing and highly refined maritime adaptation up and down the coasts of

North, Central, and South America. Geographically, the Atlantic waterway locations

of Topper and Cactus Hill could have facilitated inland access by early arrivals from

Europe’s Iberia, and Meadowcroft (and Page-Ladson) by people moving up the

Mississippi and Ohio Rivers from the Gulf of Mexico. None of these sites, or others

along the Atlantic Coast, however, suggests a maritime subsistence strategy or any

evidence for boat use, although Jodry (2006) notes later Paleoindian sites in the region

where watercraft could profitably have been used. Hebior, Schaefer, La Sena, and

Lovewell do not even have geography going for them when scrutinized as possible

manifestations of a coastal peopling, and certainly their locations and archaeological

assemblages belie a coastal origin and maritime adaptation.

In fact, all eight of the problematic sites (problematic for my peopling model that

is), present a relatively consistent picture of a terrestrial hunting adaptation in North

America starting during the LGM, when I have suggested people also occupied

western Beringia. La Sena and Lovewell yielded direct evidence for human

butchering of proboscideans during the LGM, as did Hebior, Schaefer, and Page-

Ladson some 5,000 years later (but more than a millennium prior to Clovis).

Topper, Cactus Hill, and Meadowcroft all revealed tool kits dating to the LGM or

shortly thereafter that included small prismatic blades and true microblades, blade

cores, and in a later but still pre-Clovis occupation at Meadowcroft, bifacial

projectile points. Page-Ladson, too, yielded chipped stone artifacts, albeit more

expedient unifacial tools and debitage, together with the mastodon remains.

356 J Archaeol Res (2011) 19:327–375

123

Page 31: fulltext (2)

Although some of the ‘‘problem’’ sites offer more clues to the subsistence strategy

of early occupants than others, none suggests anything other than a terrestrial

hunting and gathering lifeway—and a chipped stone and bone tool kit—that is

entirely consistent with (a) slightly later interior eastern Beringian sites like Swan

Point, (b) later-still Clovis, and (c) known pre-LGM (e.g., Yana RHS), LGM (e.g.,

Ikhine 2), and immediately post-LGM sites in western Beringia (e.g., Berelekh).

If we accept these sites at face value, we must then revisit the question of how

their occupants accessed North America. Although some disagree and actively

support peopling models that accommodate immigrants from non-Asian source

regions (e.g., Bradley and Stanford 2004; Faught 2008; Miotti 2003; Miotti and

Salemme 2003), extant evidence—archaeological, genetic, and osteological—is so

clearly consistent on this point that it compels a northeast Asian origin for all First

Americans, including those now under consideration. That leaves us to address the

route and mechanism for the arrival of what may be the New World’s earliest

residents. Leaving intact my logic for viewing well-documented early coastal

residents and Clovis as representing different pulses of migration from Western

Beringia, the former by boat and sea and the latter by land, I read the evidence from

these potentially earliest New World sites as most consistent with an interior IFC

entry at some point or points prior to 14 k–13 k cal. B.P. Such an entry would

bolster arguments offered recently by some that Clovis per se may have originated

not in the north as traditionally thought, but in the southeast, with the technology

spreading north, east, and west through extant populations (e.g., Anderson 2004;

Anderson and Faught 2000; Beck and Jones 2010; Bradley and Stanford 2004;

Roper and Wygal 2002; cf. Steele 2009; Steele and Politis 2009).

Anderson and Gillam’s (2000) least-cost colonization simulations showed that an

IFC entry could relatively neatly account for the locations, if not so neatly the

chronology, of the sites at issue. They modeled a primary pathway that led

immigrants from the mouth of the IFC to the Missouri River, then the Mississippi

and on to the Gulf of Mexico, and from there north along the east coast. This path

would have allowed people to access Page-Ladson, Topper, and Cactus Hill rather

directly, and with a detour east along the Ohio River, Meadowcroft. Anderson and

Gillam (2000) also posited secondary branches that would have facilitated access to

the Central Plains generally, and Lovewell and La Sena specifically, via westward

travel along the Platte River, a tributary of the Missouri that flows through south-

central Nebraska. A trip overland across the Dakotas, Minnesota or Iowa, and

Wisconsin—another secondary least-cost pathway (Anderson and Gillam 2000)—

leads to the Great Lakes region and Schaefer and Hebior.

The problem with both the application of Anderson and Gillam’s (2000)

simulation data for an IFC entry to this particular suite of early sites and the

hypothesis of an IFC entry for them generally is rooted in timing. A significant body

of paleoecological data, cited earlier, supports closure of the (or ‘‘an’’) IFC between

30,000–24,000 and 14,000–13,000 years ago (and of the northern Pacific Coast

before 16,000–15,000 years ago). Embracing a northeast Asian origin, which

myriad evidence indicates we must, leaves a serious hurdle: getting people south of

the Laurentide and Cordilleran ice sheets by peak LGM time so they could leave

behind the late LGM-era La Sena and Lovewell sites of the High Plains, and

J Archaeol Res (2011) 19:327–375 357

123

Page 32: fulltext (2)

Topper, Cactus Hill, and Meadowcroft in the eastern U.S. To do this on the basis of

current paleoenvironmental data requires, most parsimoniously, a return to c.

30,000–24,000 years ago, the last time the IFC could have served as a conduit for

human migration from eastern Beringia to the continental U.S.

A scenario of pre-LGM movement through the IFC is not inherently unsupport-

able on the basis of current evidence as is, for example, a non-Asian origin for New

World colonization. The latest linguistic-based peopling treatise in fact supports it

(Nichols 2008). Archaeologically, humans unequivocally occupied the 32 k cal.

B.P. western Beringian Yana RHS with its rhino-horn rods and bifacial chipped

stone technology. Morlan (2003) also has argued, although not always to a receptive

audience (e.g., Bever 2001), that Eastern Beringians knapped proboscidean bone in

the vicinity of central Alaska’s Bluefish Caves as early as c. 28 k cal. B.P., and

maybe in the Old Crow Basin for many millennia before that (but see Dixon [1999]

and Fiedel [2000] for examples of dissenting views). On the one hand, the

assemblages of the Yana RHS, Bluefish Caves, the Old Crow region, and the very

earliest New World site candidates, La Sena and Lovewell, are strikingly similar

with their bone rods and apparent mammoth-bone-flaking technologies. On the

other hand, thus modeling a pre-LGM IFC entry to the New World leaves a

minimum gap of six millennia between the youngest (and again, controversial) pre-

LGM eastern Beringian sites and the oldest candidates from the lower 48 states. It

also suggests an older peopling time frame than many recent molecular studies

support.

Another chronological problem arises if we subscribe to a pre-LGM North

American colonization process that even remotely resembles that simulated by

Anderson and Gillam (2000) for an IFC entry. Their proposed primary routing of

migration south and throughout the continent suggests that sites along or reasonably

associated with the Gulf and East Coasts, for example, should show progressively

younger occupations as immigrants made their way down the Missouri and

Mississippi, and then east and north along the coasts. If the sites under consideration

here are all part of the same initial, pre-LGM IFC peopling, we might expect Page-

Ladson to be oldest, then Topper, then Cactus Hill. But precisely the reverse is true.

We might also expect La Sena and Lovewell to be among the younger sites, given

that least-cost analysis pinpoints movement along the major north-south rivers

systems of the interior continent as optimum. On the other hand, Anderson and

Gillam (2000, p. 47) point out that a colonizing population had, by definition, no

awareness of what lay ahead, so a reasonably appealing Platte River route could

have beckoned to some First Americans to move west.

Conclusions and future research directions

In 2010, and in the spirit of writing the best narrative possible given current

evidence, I conclude that genetic, osteological, and archaeological evidence from

the Old World and the New World can be marshaled to support a peopling of the

Americas that began 16 k–15 k cal. B.P. and ensued in two major pulses. Both

originated most immediately during the LGM in western Beringia, and before that

358 J Archaeol Res (2011) 19:327–375

123

Page 33: fulltext (2)

in south-central and southeastern Siberia. The first major migratory episode

occurred by watercraft and followed the northern Pacific Rim south along the coast

of Alaska and North America. At the Isthmus of Tehuantepec, coastal migration

proceeded south to South America and perhaps also north along the coast of the

Gulf of Mexico. At the Isthmus of Panama, it again split such that people moved

south along the east and west coasts of South America. The second major migration

ensued when the Cordilleran and Laurentide ice sheets receded, opening an IFC that

people could pass through by 14 k–13 k cal. B.P. The terrestrial hunter-gatherers

who passed through the IFC developed the Clovis tool kit and used it to settle

throughout North and Central America by the end of the Pleistocene.

Tantalizing evidence suggests that the complete picture of the peopling process

may involve an even earlier wave of colonization, one that proceeded on foot

through a pre-LGM IFC by Upper Paleolithic terrestrial hunter-gatherers, and that

could have spawned Clovis technology in the southeast that radiated north, east, and

west. I do not embrace it wholesale, not so much because I reject the North

American sites that collectively support it but because we cannot now plug a

minimum (and maybe much greater than) 6,000-year time gap between eastern

Beringian occupation and sites in the continental U.S, and to a lesser degree because

the pre-LGM time frame conflicts with the results of many recent genetic estimates

for colonization. The time-gap issue could be resolved in one of two ways: by

identifying alternative routes south between, around, or over the ice sheets during

the LGM or by locating North American sites that date conclusively to c. 28 k–

22 k cal. B.P. Because those dates represent controversial sites (Bluefish Caves in

the former instance and La Sena/Lovewell in the latter), it also would be helpful to

identify additional localities dating to those same time frames in eastern Beringia

and North America, respectively. Should either body of evidence emerge, I do not

foresee a need to radically overhaul the rest of the scenario I have proposed. The

source region would remain the same, as would the genetic fingerprint of the source

population.

Although I have alluded to some of these in the preceding section, I end with a

few explicitly stated research directions that could produce evidence to evaluate the

scenario I have proposed. First, we need to intensify the search for LGM sites in

western Beringia, both inland and along LGM shorelines. We have the ability to use

geologic, paleoecological, and other data to create strong predictive models for

possible site locations, and we need to do so and then follow through with sustained

ground-truthing efforts. Similarly, we must intensify the search for now-submerged

early sites along the west coast of North and South America, and also inland sites

located along major river corridors. A few are doing all of this, but collectively not

with the intensity necessary—or in the array of places necessary—to rigorously

evaluate the scenario I have proposed or any coastal migration scenario.

Coastal-route advocates and detractors agree that unless and until we find very

ancient sites under Pacific coastal waters, the case for a coastal peopling by

watercraft will remain difficult to support. However, I sense that broadly speaking

(again, a few progressive researchers not withstanding), we are paralyzed when it

comes to mounting full-blown efforts to learn what is on the sea floor. Yes indeed, it

is a big ocean and a mighty long coastline, and yes indeed, it will be hugely

J Archaeol Res (2011) 19:327–375 359

123

Page 34: fulltext (2)

expensive and logistically challenging to find and document sites that are now

underwater—or to explore enough sea-bottom to demonstrate they are not there. Yet

Fedje and Josenhans (2000) have shown that reconstructing underwater landscapes

is possible, and Faught (e.g., 2004) has demonstrated that early underwater sites can

be found, can be well preserved, and can be excavated. Predictive modeling can

point the way to the richest Late Pleistocene coastal ‘‘oases’’ (Dixon 1999), and

methods will have to be tailored to a given underwater landscape (i.e., some may be

penetrable by human divers, others may require submersibles). But the time has

come to go ‘‘all in’’ and turn hypothetical arguments into well-supported versions.

There are many other directions we can go: continuing to evaluate the Clovis

radiocarbon record to assess whether the age cline trends north to south, as I have

argued, or south to north as some others have recently proposed (e.g., Beck and

Jones 2010); characterizing the chipped stone technologies of early sites in major

western river corridors to assess whether and how these changed as coastal foragers

moved inland; predicting which river corridors may have beckoned most alluringly

to coastal hunter-gatherers of the terminal Pleistocene; exploring, perhaps through

judicious use of ethnographic analogy, what may have transpired—and what sort of

signature it may have left behind—if and when coastal foragers moving east

encountered Clovis-era foragers moving west; bringing to bear case studies and

invoking computer modeling to explore circumstances under which foragers change

subsistence strategies (as from coastal foraging to terrestrial); and of course,

continuing the search for sources of very ancient DNA in the New World and

northeastern Asia to confirm, expand, or otherwise refine our understanding of the

full array of the haplogroups of First Americans and their ancestors. The next

decade will see the pursuit of these and other research directions, and the only sure

bet is that at least a few pending peopling finds will be of the ‘‘wow—I never saw

that coming’’ variety.

Acknowledgments First, I thank Gary Feinman for inviting me to write this article and for his

encouragement and helpful suggestions throughout the publication process. I also thank David Anderson,

Charlotte Beck, and four anonymous reviewers for their concrete and constructive feedback, which was

offered, to a person, without gratuitous nastiness or toxic rhetoric. I also gratefully acknowledge my many

‘‘Peopling of the New World’’ students who, through the years, have challenged my thinking on the

subject every bit as much as my colleagues in the field have done. Finally, I thank my Ph.D. mentor,

Vance Haynes, for inspiring and encouraging me to think openly about the peopling of the New World

and for setting a model for careful thinking and civil scholarship that I always have and always will try to

emulate.

References cited

Abromova, Z. A. (1989). Paleolit Severnoi Azii (Paleolithic of northern Asia). In Boriskovsky, P. I. (ed.),

Paleolit Kavkaza I Severnoi Azii, Nauka, Leningrad, pp. 145–243.

Achilli, A., Perego, U. A., Bravi, C. M., Coble, M. D., Kong, Q. P., Woodward, S. R., Salas, A., Torroni,

A., and Bandelt, H. J. (2008). The phylogeny of the four pan-American mtDNA haplogroups:

Implications for evolutionary and disease studies. PLoS ONE 3:e1764.

Acosta, J de. (2002 [1604]). The Natural and Moral History of the Indies, edited by Mangan, J. E., and

translated by Lopez-Morillas, F., Duke University Press, Durham, NC.

360 J Archaeol Res (2011) 19:327–375

123

Page 35: fulltext (2)

Adovasio, J., and Page, J. (2003). The First Americans: In Pursuit of Archaeology’s Greatest Mystery,

Modern Library, New York.

Adovasio, J. M., Pedler, D., Donahue, J., and Stuckenrath, R. (1999). No vestige of a beginning nor

prospect for an end: Two decades of debate on Meadowcroft Rockshelter. In Bonnichsen, R., and

Turmire, K. L. (eds.), Ice Age Peoples of North America: Environments, Origins, and Adaptations ofthe First Americans, Oregon State University Press, Corvallis, pp. 416–431.

Agenbroad, L. D., Johnson, J. R., Morris, D., and Stafford, Jr., T. W. (2005). Mammoths and humans as

Late Pleistocene contemporaries on Santa Rosa Island. In Garcelon, D., and Schwemm, C. (eds.),

Proceedings of the 6th California Islands Symposium. National Park Service Technical Publication

CHIS-05-01, Institute for Wildlife Studies, Arcata, CA, pp. 3–7.

Anderson, D. G. (2004). Paleoindian occupations in the southeastern United States. In Lepper, B. T., and

Bonnichsen, R. (eds.), New Perspectives on the First Americans, Texas A&M University Press,

College Station, pp. 119–128.

Anderson, D. G. (2005). Pleistocene human occupation of the Southeastern United States: Research

directions for the early 21st century. In Bonnichsen, R., Lepper, B. T., Stanford, D., and Waters, M.

R. (eds.), Paleoamerican Origins: Beyond Clovis, Texas A&M University Press, College Station,

pp. 29–43.

Anderson, D. G., and Faught, M. K. (2000). Paleoindian artefact distributions: Evidence and implications.

Antiquity 74: 507–513.

Anderson, D. G., and Gillam, J. C. (2000). Paleoindian colonization of the Americas: Implications from

an examination of physiography, demography, and artifact distribution. American Antiquity 65:43–66.

Anderson, D. G., and Gillam, J. C. (2001). Paleoindian interaction and mating networks: Reply to Moore

and Moseley. American Antiquity 66: 530–535.

Anderson, D. G., Miller, D. S., Yerka, S. J., and Faught, M. K. (2005). Paleoindian database of the

Americas: 2005 status report. Current Research in the Pleistocene 22: 91–92.

Anderson, D. G., Miller, D. S., Yerka, S. J., Gillam, J. C., Johanson, E. N., Anderson, D. T., Goodyear, A. C.,

and Smallwood, A. M. (2010). PIDBA (Paleoindian database of the Americas) 2010: Current status and

findings. Archaeology of Eastern North America 38: 63–90.

Armour, J. A. L., Anttinen, T., May, C. A., Vega, E. E., and Sajantila, A. (1996). Minisatellite diversity

supports a recent African origin for modern humans. Nature Genetics 13: 154–160.

Arnold, T. G. (2002). Radiocarbon dates from the ice-free corridor. Radiocarbon 44: 437–454.

Atkinson, Q. D., Gray, R. D., and Drummond, A. J. (2008). MtDNA variation predicts population size in

humans and reveals a major Southern Asian chapter in human prehistory. Molecular Biology andEvolution 25: 468–474.

Auerbach, B. M. (2008). Human Skeletal Variation in the New World During the Holocene: Effects ofClimate and Subsistence Across Geography and Time, Part 1, ProQuest Information and Learning

Company, Ann Arbor, MI.

Bandelt, H. J., Herrnstadt, C., Yao, Y. G., Kong, Q. P., Kivisild, T., Rengo, C., Scozzari, R., Richards, M.,

Villems, R., Macaulay, V., Howell, N., Torroni, A., and Zhang, Y. P. (2003). Identification of Native

American founder mtDNAs through the analysis of complete mtDNA sequences: Some caveats.

Annals of Human Genetics 67: 512–524.

Barnes, I., Shapiro, B., Lister, A., Kuznetova, T., Sher, A., Guthrie, D., and Thomas, M. G. (2007).

Genetic structure and extinction of the Woolly Mammoth, Mammuthus primigenius. CurrentBiology 17: 1072–1075.

Battilana, J., Fagundes, N. J. R., Heller, A. H., Goldani, A., Freitas, L. B., Tarazona-Santos, E.,

Munkhbat, B., Munkhtuvshin, N., Krylov, M., Benevolenskaia, L., Arnett, F. C., Batzer, M. A.,

Deininger, P. L., Salzano, F. M., and Bonatto, S. L. (2006). Alu insertion polymorphisms in Native

Americans and related Asian populations. Annals of Human Biology 33: 142–160.

Battilana, J., Cardoso-Silva, L., Barrantes, R., Hill, K., Hurtado, A. M., Salzano, F. M., and Bonatto, S. L.

(2007). Molecular variability of the 16p13.3 region in Amerindians and its anthropological

significance. Annals of Human Genetics 71: 64–76.

Beaton, J. M. (1991). Colonizing continents: Some problems from Australia and the Americas. In

Dillehay, T. D., and Meltzer, D. J. (eds.), The First Americans: Search and Research, CRC Press,

Boca Raton, FL, pp. 209–230.

Beck, C., and Jones, G. T. (2010). Clovis and Western Stemmed: Population migration and the meeting of

two technologies in the intermountain west. American Antiquity 75: 81–116.

J Archaeol Res (2011) 19:327–375 361

123

Page 36: fulltext (2)

Bever, M. R. (2001). An overview of Alaskan Late Pleistocene archaeology: Historical themes and

current perspectives. Journal of World Prehistory 15: 125–191.

Bever, M. R. (2006). Too little, too late? The radiocarbon chronology of Alaska and the peopling of the

New World. American Antiquity 71: 959–620.

Borrero, L. (1999). Human dispersal and climatic conditions during the late Pleistocene times in Fuego-

Patagonia. Quaternary International 53–54: 93–99.

Bortolini, M. C., Salzano, F. M., Thomas, M. G., Stuart, S., Nasanen, S. P., Bau, C. H., Hutz, M. H.,

Layrisse, Z., Petzl-Erler, M. L., Tsuneto, L. T., Hill, K., Hurtado, A. M., Castro-de-Guerra, D.,

Torres, M. M., Groot, H., Michalski, R., Nymadawa, P., Bedoya, G., Bradman, N., Labuda, D., and

Ruiz-Linares, A. (2003). Y-chromosome evidence for differing ancient demographic histories in the

Americas. American Journal of Human Genetics 73: 524–539.

Brace, L. C., Nelson, A. R., Seguchi, N., Oe, O., Sering, L., Qifeng, P., Yongyi, L., and Tumen, D.

(2001). Old World sources of the first New World human inhabitants: A comparative craniofacial

view. Proceedings of the National Academy of Sciences 98: 10017–10022.

Bradley, B., and Stanford, D. (2004). The North Atlantic ice-edge corridor: A possible Paleolithic route to

the New World. World Archaeology 36: 459–478.

Brubaker, L. B., Anderson, P. M., Edwards, M. E., and Lozhkin, A. V. (2005). Beringia as a glacial

refugium for boreal trees and shrubs: New perspectives from mapped pollen data. Journal ofBiogeography 32: 833–848.

Bryan, A. L., and Gruhn, R. (2003). Some difficulties in modeling the original peopling of the Americas.

Quaternary International 109–110: 175–179.

Buchanan, B., and Collard, M. (2007). Investigations of the peopling of North America through cladistic

analyses of Early Paleoindian projectile points. Journal of Anthropological Archaeology 26:366–393.

Burns, J. A. (1996). Vertebrate paleontology and the alleged ice-free corridor: The meat of the matter.

Quaternary International 32: 107–112.

Byers, D. A., and Ugan, A. (2005). Should we expect large game specialization in the late Pleistocene?

An optimal foraging perspective on early Paleoindian prey choice. Journal of ArchaeologicalScience 32: 1624–1640.

Cannon, A. (2000). Settlement and sea-levels on the central coast of British Columbia: Evidence from

shell midden cores. American Antiquity 65: 67–77.

Cannon, M. D., and Meltzer, D. J. (2004). Early Paleoindian foraging: Examining the faunal evidence for

large mammal specialization and regional variability in prey choice. Quaternary Science Reviews23: 1955–1987.

Carlson, R. L. (1998). Coastal British Columbia in the light of North Pacific maritime adaptations. ArcticAnthropology 35: 23–35.

Chatters, J. C. (2000). The recovery and first analysis of an Early Holocene human skeleton from

Kennewick, Washington. American Antiquity 65: 291–316.

Clague, J. J., and James, T. S. (2002). History and isostatic effects of the last ice sheet in southern British

Columbia. Quaternary Science Reviews 21: 71–87.

Clark, P. U., and Mix, A. C. (2002). Ice sheets and sea level of the last glacial maximum. QuaternaryScience Reviews 21: 1–7.

Comas, D., Plaza, S., Wells, R. S., Yuldaseva, N., Lao, O., Calafell, F., and Bertranpetit, J. (2004).

Admixture, migrations, and dispersals in Central Asia: Evidence from maternal DNA lineages.

European Journal of Human Genetics 12: 495–504.

Cunningham, D., and Jantz, R. L. (2003). The morphometric relationship of Upper Cave 101 and 103 to

modern Homo sapiens. Journal of Human Evolution 45: 1–18.

Davis, L. G., and Schweger, C. E. (2004). Geoarchaeological context of Late Pleistocene and Early

Holocene occupation at the Cooper’s Ferry site, western Idaho, USA. Geoarchaeology 19: 685–704.

Davis, L. G., Punke, M. L., Hall, R. L., Fillmore, M., and Willis, S. C. (2002–2004). A Late Pleistocene

occupation on the southern coast of Oregon. Journal of Field Archaeology 29: 7–16.

deFrance, S. D., Keefer, D. K., Richardson, J. B., and Alvarez, A. U. (2001). Late Paleo-Indian coastal

foragers: Specialized extractive behavior at Quebrada Tacahuay, Peru. Latin American Antiquity 12:413–426.

deFrance, S. D., Grayson, N., and Wise, K. (2009). Documenting 12,000 years of coastal occupation on

the Osmore littoral, Peru. Journal of Field Archaeology 34: 227–246.

362 J Archaeol Res (2011) 19:327–375

123

Page 37: fulltext (2)

Derbeneva, O. A., Sukernik, R. I., Volodko, N. V., Hosseini, S. H., Lott, M. T., and Wallace, D. C.

(2002). Analysis of mitochondrial DNA diversity in the Aleuts of the Commander Islands and its

implications for the genetic history of Beringia. American Journal of Human Genetics 71: 415–421.

Derenko, M. V., Gryzbowski, T., Malyarchuk, B. A., Czarny, J., Miscicka-Sliwka, D., and Zakharov, I. A.

(2001). The presence of mitochondrial haplogroup X in Altaians from South Siberia. AmericanJournal of Human Genetics 69: 237–241.

Derenko, M., Malyaruchuk, B., Grzybowski, T., Denisova, G., Dambueva, I., Perkova, M., Dorhzu, C.,

Luzina, F., Lee, H. K., Vanecek, T., Villems, R., and Zakhorov, I. (2007). Phylogeographic analysis

of mitochondrial DNA in northern Asian populations. American Journal of Human Genetics 81:1025–1041.

Dillehay, T. D. (ed.). (1999). Monte Verde: A Late Pleistocene Settlement in Chile, Volume 2: TheArchaeological Context and Interpretation, Smithsonian Institution Press, Washington, DC.

Dillehay, T. D. (2009). Probing deeper into first American studies. Proceedings of the National Academyof Sciences 106: 971–978.

Dillehay, T. D., Ramırez, C., Pino, M., Collins, M. B., Rossen, J., and Pino-Navarro, J. D. (2008). Monte

Verde: Seaweed, food, medicine, and the peopling of South America. Science 320: 784–786.

Dixon, E. J. (1999). Bones, boats, and bison: Archaeology and the first colonization of Western NorthAmerica, University of New Mexico Press, Albuquerque.

Dixon, E. J. (2001). Human colonization of the Americas: Timing, technology and process. QuaternaryScience Review 20: 277–299.

Dolukhanov, P. M., Shukurov, A. M., Tarasov, P. E., and Zaitseva, G. I. (2002). Colonization of northern

Eurasia by modern humans: Radiocarbon chronology and environment. Journal of ArchaeologicalScience 29: 593–606.

Dornelles, C. L., Bonatto, S. L., de Freitas, L. B., and Salzano, F. M. (2005). Is haplogroup X present in

extant South American Indians? American Journal of Physical Anthropology 127: 439–448.

Dumond, D. E. (2001). The archaeology of Eastern Beringia: Some contrasts and connections. ArcticAnthropology 38: 196–205.

Dunbar, J. S. (2006a). Paleoindian archeology. In Webb, S. D. (ed.), First Floridians and LastMastodons: The Page-Ladson site in the Aucilla River, Springer, Dordrecht, The Netherlands,

pp. 403–435.

Dunbar, J. S. (2006b). Pleistocene-Early Holocene climate change: Chronostratigraphy and geoclimate of

the southeast US. In Webb, S. D. (ed.), First Floridians and Last Mastadons: The Page-Ladson sitein the Aucilla River, Springer, Dordrecht, The Netherlands, pp. 103–155.

Dyke, A. S. (2004). An outline of North American deglaciation with emphasis on central and northern

Canada. In Ehlers, J., and Gibbard, P. L. (eds.), Quaternary Glaciations: Extent and Chronology,

Part II, Elsevier Press, Amsterdam, pp. 373–424.

Elston, R. G., and Zeanah, D. W. (2002). Thinking outside the box: A new perspective on diet breadth and

sexual division of labor in the Prearchaic Great Basin. World Archaeology 34: 103–130.

Erickson, D. L., Smith, B. D., Clarke, A. C., Sandweiss, D. H., and Tuross, N. (2005). An Asian origin for

a 10,000-year-old domesticated plant in the Americas. Proceedings of the National Academy ofSciences 102: 18315–18320.

Erlandson, J. M. (2002). Anatomically modern humans, maritime voyaging, and the Pleistocene

colonization of the Americas. In Jablonski, N. G. (ed.), The First Americans: The PleistocenePeopling of the New World, Memoirs No. 27, California Academy of Sciences, San Francisco,

pp. 59–92.

Erlandson, J. M., Braje, T. J., Rick, T. C., and Peterson, J. (2005). Beads, bifaces, and boats: An early

maritime adaptation on the south coast of San Miguel Island, California. American Anthropologist107: 677–683.

Erlandson, J. M., Graham, M. H., Bourque, B. J., Corbett, D., Estes, J. A., and Steneck, R. S. (2007). The

kelp highway hypothesis: Marine ecology, the coastal migration theory, and the peopling of the

Americas. Journal of Island and Coastal Archaeology 2: 161–174.

Erlandson, J. M., Moss, M. L., and des Lauriers, M. (2008). Life on the edge: Early maritime cultures of

the Pacific Coast of North America. Quaternary Science Reviews 27: 2232–2245.

Eshleman, J. A., Malhi, R. S., and Smith, D. G. (2003). Mitochondrial DNA studies of Native Americans:

Conceptions and misconceptions of the population prehistory of the Americas. EvolutionaryAnthropology 12: 7–18.

Fagan, B. (2004). The House of the sea: An essay on the antiquity of planked canoes in Southern

California. American Antiquity 69: 7–16.

J Archaeol Res (2011) 19:327–375 363

123

Page 38: fulltext (2)

Fagundes, N. J. R., Kanitz, R., Eckert, R., Valls, A. C. S., Bogo, M. R., Salzano, F. M., Smith, D. G.,

Silva, Jr., W. A., Zago, M. A., Ribeiro-dos-Santos, A. K., Santos, S. E. B., Petzl-Erler, M. L., and

Bonatto, S. L. (2008). Mitochondrial population genomics supports a single Pre-Clovis origin with a

costal route for the peopling of the Americas. The American Journal of Human Genetics 82:583–592.

Falk, T. (2009). Wisconsin dig seeks to confirm Pre-Clovis Americans. Science 305: 590.

Faught, M. K. (2002–2004). Submerged Paleoindian and Archaic sites of the Big Bend, Florida. Journalof Field Archaeology 29: 273–290.

Faught, M. K. (2004). The underwater archaeology of paleolandscapes, Apalachee Bay, Florida.

American Antiquity 69: 275–289.

Faught, M. K. (2006). Paleoindian archeology in Florida and Panama. In Morrow, J., and Gnecca, C.

(eds.), Paleoindian Archeology: A Hemispheric Perspective, University Press of Florida, Gaines-

ville, pp. 164–183.

Faught, M. K. (2008). Archaeological roots of human diversity in the New World: A compilation of

accurate and precise radiocarbon dates from earliest sites. American Antiquity 73: 670–698.

Feathers, J. K., Rhodes, E. J., Huot, S., and Macavoy, J. M. (2006). Luminescence dating of sand deposits

related to late Pleistocene human occupation at the Cactus Hill Site, Virginia, USA. QuaternaryGeochronology 1: 167–187.

Fedje, D. W., and Josenhans, H. (2000). Drowned forests and archaeology on the continental shelf of

British Columbia, Canada. Geology 28: 99–102.

Fedje, D. W., and Southon, J. (2003). A post-glacial record of 14C reservoir ages for the British Columbia

coast. Canadian Journal of Archaeology 27: 95–111.

Fedje, D. W., Mackie, Q., Dixon, E. J., and Heaton, T. H. (2004). Late Wisconsin environments and

archaeological visibility on the northern Northwest Coast. In Madsen, D. B. (ed.), Entering America:Northeast Asia and Beringia before the Last Glacial Maximum, University of Utah Press, Salt Lake

City, pp. 97–139.

Fiedel, S. J. (1999). Older than we thought: Implications of corrected dates for Paleoindians. AmericanAntiquity 64: 95–115.

Fiedel, S. J. (2000). The peopling of the New World: Present evidence, new theories, and future

directions. Journal of Archaeological Research 8: 39–103.

Fiedel, S. J. (2002). Initial human colonization of the Americas: An overview of the issues and evidence.

Radiocarbon 44: 407–436.

Fiedel, S. J. (2004). Kennewick Follies: ‘‘New’’ theories about the peopling of the Americas. Journal ofAnthropological Research 60: 75–110.

Fiedel, S. J. (2005). Man’s best friend—mammoth’s worst enemy? A speculative essay on the role of

dogs in Paleoindian colonization and megafaunal extinction. World Archaeology 37: 11–25.

Field, J. S., and Lahr, M. M. (2006). Assessment of the southern dispersal: GIS-based analyses of

potential routes at oxygen isotope stage 4. Journal of World Prehistory 19: 1–45.

Field, J. S., Petraglia, M. D., and Lahr, M. M. (2007). The southern dispersal hypothesis and the South

Asian archaeological record: Examination of dispersal routes through GIS analysis. Journal ofAnthropological Archaeology 26: 88–108.

Firestone, R. B., West, A., Kennett, J. P., Becker, L., Bunch, T. E., Revay, Z. S., Schultz, P. H., Belgya,

T., Kennett, D. J., Erlandson, J. M., Dickenson, O. J., Goodyear, A. C., Harris, R. S., Howard, G. A.,

Kloosterman, J. B., Lechler, P., Mayewski, P. A., Montgomery, J., Poreda, R., Darrah, T., Que Hee,

S. S., Smith, A. R., Stitch, A., Topping, W., Wittke, J. H., and Wolbach, W. S. (2007). Evidence for

an extraterrestrial impact 12,900 years ago that contributed to the megafaunal extinctions and the

Younger Dryas cooling. Proceedings of the National Academy of Sciences 104: 16016–16021.

Fitzgerald, R. T., and Jones, T. L. (2003). On the weight of the evidence from Cross Creek: A reply to

Turner. American Antiquity 68: 396–399.

Fix, A. G. (2005). Rapid deployment of the five founding Amerind mtDNA haplogroups via coastal and

riverine colonization. American Journal of Physical Anthropology 128: 430–436.

Fladmark, K. R. (1979). Routes: Alternate migration corridors for early man in North America. AmericanAntiquity 44: 55–69.

Fladmark, K. R. (1986). Getting one’s Berings. Natural History 95: 8–17.

Forster, P. (2004). Ice ages and the mitochondrial DNA chronology of human dispersals: A review.

Philosophical Transactions: Biological Sciences 359: 255–264.

Forster, P., Torroni, A., Renfrew, C., and Rohl, A. (2001). Phylogenetic star contraction applied to Asian

and Papuan mtDNA evolution. Molecular Biology and Evolution 18: 1864–1881.

364 J Archaeol Res (2011) 19:327–375

123

Page 39: fulltext (2)

Gilbert, M. T. P., Kivisild, T., Grønnow, B., Andersen, P. K., Metspalu, E., Reidla, M., Tamm, E.,

Axelsson, E., Gotherstrom, A., Campos, P. F., Rasmussen, M., Metspalu, M., Higham, T. F. G.,

Schwenninger, J-L., Nathan, R., De Hoog, C-J., Koch, A., Nukaaraq Møller, L., Andreasen, C.,

Meldgaard, M., Villems, R., Bendixen, C., and Willerslev, E. (2008a). Paleo-Eskimo mtDNA

genome reveals matrilineal discontinuity in Greenland. Science 320: 1787–1789.

Gilbert, M. T. P., Jenkins, D. L., Gotherstrom, A., Naveran, N., Sanchez, J. J., Hofreiter, M., Thomsen, P.

F., Binladen, J., Higham, T. F. G., Yohe II, R. M., Parr, R., Cummings, L. S., and Willerslev, E.

(2008b). DNA from Pre-Clovis human coprolites in Oregon, North America. Science 320: 786–789.

Gilbert, M. T. P., Jenkins, D. L., Higham, T. F. G., Rasmussen, M., Malmstrom, H., Svensson, E. M.,

Sanchez, J. J., Cummings, L. S., Yohe II, R. M., Hofreiter, M., Gotherstrom, A., and Willerslev, E.

(2009). Response to Comment by Poinar et al. on ‘‘DNA from Pre-Clovis human coprolites in

Oregon, North America.’’ Science 325: 148b.

Goebel, T. (1999). Pleistocene human colonization of Siberia and peopling of the Americas: An

ecological approach. Evolutionary Anthropology 8: 208–227.

Goebel, T. (2002). The ‘‘microblade adaptation’’ and recolonization of Siberia during the late Upper

Pleistocene. In Elston, R. G., and Kuhn, S. L. (eds.), Thinking Small: Global Perspectives onMicrolithization, American Anthropological Association, Arlington, VA, pp. 117–131.

Goebel, T. (2007). The missing years for modern humans. Science 315: 194–196.

Goebel, T., Waters, M. R., and Dikova, M. (2003). The archaeology of Ushki Lake, Kamchatka, and the

Pleistocene peopling of the Americas. American Antiquity 301: 501–505.

Goebel, T., Waters, M. R., and O’Rourke, D. H. (2008). The Late Pleistocene dispersal of modern

humans in the Americas. Science 319: 1497–1502.

Goldberg, P., Berna, F., and Macphail, R. I. (2009). Comment on ‘‘DNA from Pre-Clovis human

coprolites in Oregon, North America.’’ Science 325: 148c.

Gonzalez-Jose, R., Neves, W., Lahr, M. M., Gonzalez, S., Pucciarelli, H., Martınez, M. H., and Correal,

G. (2005). Late Pleistocene/Holocene craniofacial morphology in Mesoamerican Paleoindians:

Implications for the peopling of the New World. American Journal of Physical Anthropology 128:772–780.

Goodyear, A. C. (2006). Evidence for pre-Clovis sites in the eastern United States. In Bonnichsen, R.,

Lepper, B. T., Stanford, D., and Waters, M. R. (eds.), Paleoamerican Origins: Beyond Clovis,

Center for the Study of the First Americans, Texas A&M University, College Station, pp. 103–112.

Goodyear, A. C., and Steffy, K. (2003). Evidence of a Clovis occupation at the Topper site, 28AL23,

Allendale County, South Carolina. Current Research in the Pleistocene 20: 23–25.

Grayson, D. K., and Meltzer, D. J. (2003). A requiem for North American overkill. Journal ofArchaeological Science 30: 585–593.

Green, T. J., Cochran, B., Fenton, T. W., Woods, J. C., Titmus, G. L., Tiezen, L., Davis, M. A., and

Miller, S. J. (1998). The Buhl burial: A Paleoindian woman from southern Idaho. AmericanAntiquity 63: 437–456.

Gruhn, R. (2006). Reconstructing prehistoric population movements: Seeking congruence in genetics,

linguistics, and archaeology. Reviews in Anthropology 35: 345–372.

Hamilton, M. J., and Buchanan, B. (2007). Spatial gradients in Clovis-age radiocarbon dates across North

America suggest rapid colonization from the north. Proceedings of the National Academy ofSciences 104: 15625–15630.

Haynes, Jr., C. V. (1964). Fluted projectile points: Their age and dispersion. Science 145: 1408–1413.

Haynes, G. (2002). The Early Settlement of North America, Cambridge University Press, Cambridge.

Haynes, G., Anderson, D. G., Ferring, C. R., Fiedel, S. J., Grayson, D. K., Haynes, Jr., C. V., Holliday, V. T.,

Huckell, B. B., Kornfeld, M., Meltzer, D. J., Morrow, J., Surovell, T., Waguespack, N. M., Wigand, P.,

and Yohe II, R. M. (2007). Comment on ‘‘Redefining the age of Clovis: Implications for the peopling of

the Americas.’’ Science 317: 320b.

Hazelwood, L., and Steele, J. (2003). Colonizing new landscapes: Archaeological detectibility of the first

phase. In Rockman, M., and Steele, J. (eds.), Colonization of Unfamiliar Landscapes: TheArchaeology of Adaptation, Routledge, New York, pp. 203–219.

Hazelwood, L., and Steele, J. (2004). Spatial dynamics of human dispersals: Constraints on modeling and

archaeological validation. Journal of Archaeological Science 31: 669–679.

Hey, J. (2005). On the number of New World founders: A population genetic portrait of the peopling of

the Americas. PLoS Biology 3: 965–975.

Ho, S. Y., and Endicott, P. (2008). The crucial role of calibration in molecular date estimates for the

peopling of the Americas. American Journal of Human Genetics 83: 142–147.

J Archaeol Res (2011) 19:327–375 365

123

Page 40: fulltext (2)

Ho, S. Y. W., and Larson, G. (2006). Molecular clocks: When times are a-changin’ Trends in Genetics22: 79–83.

Hoffecker, J. F. (2001). Late Pleistocene and Early Holocene sites in the Nenana River Valley, central

Alaska. Arctic Anthropology 38: 139–153.

Hoffecker, J. F. (2002). Desolate landscapes: Ice-Age Settlement in Eastern Europe, Rutgers University

Press, New Brunswick, NJ.

Hoffecker, J. F. (2005). A Prehistory of the North: Human Settlement of the Higher Latitudes, Rutgers

University Press, New Brunswick, NJ.

Hoffecker, J. F., and Elias, S. A. (2003). Environment and archaeology in Beringia. EvolutionaryAnthropology 12: 34–49.

Hoffecker, J. F., Powers, W. R., and Goebel, T. (1993). The colonization of Beringia and the peopling of

the New World. Science 259: 46–53.

Holen, S. R. (2006). Taphonomy of two last glacial maximum mammoth sites in the central Great Plains

of North America: A preliminary report on La Sena and Lovewell. Quaternary International142–143: 30–43.

Holliday, T. W. (1999). Brachial and cural indices of European late Upper Paleolithic and Mesolithic

Europeans. Journal of Human Evolution 36: 549–566.

Holmes, C. E. (2001). Tanana River Valley archaeology circa 14,000 to 9,000 B. P. Arctic Anthropology38: 154–170.

Huckleberry, G., Stein, J. K., and Goldberg, P. (2003). Determining the provenience of Kennewick Man

skeletal remains through sedimentological analyses. Journal of Archaeological Science 30:651–665.

Jackson, D., Mendez, C., Seguel, R., Maldonado, A., and Vargas, G. (2007). Initial occupation of the

Pacific Coast of Chile during Late Pleistocene times. Current Anthropology 48: 725–731.

Jantz, R. L., and Owsley, D. W. (2001). Variation among early North American crania. American Journalof Physical Anthropology 114: 146–155.

Jobling, M. A., Hurles, M. E., and Tyler-Smith, C. (2004). Human Evolutionary Genetics: Origins,Peoples and Disease, Garland Science, New York.

Jodry, M. A. (2006). Envisioning water transport technology in late-Pleistocene America. In Bonnichsen,

R., Lepper, B. T., Stanford, D., and Waters, M. R. (eds.), Paleoamerican Origins: Beyond Clovis,

Center for the Study of the First Americans, Texas A&M University, College Station, pp. 133–160.

Johnson, J. R., Stafford, Jr., T. W., Ajie, H. O., and Morris, D. P. (2002). Arlington Springs revisited. In

Brooks, D. R., Mitchell, K. C., and Chaney, H. W. (eds.), Proceedings of the 5th California IslandsSymposium, Santa Barbara Museum of Natural History, Santa Barbara, CA, pp. 541–545.

Johnson, J. R., Stafford, Jr., T. W., West, G. J., and Rockwell, T. K. (2007). Before and after the Younger

Dryas: Chronostratigraphic and paleoenvironmental research at Arlington Springs, Santa Rosa

Island, California. Eos 88: 42A-03.

Jones, G. T., Beck, C., Jones, and Hughes, R. E. (2003). Lithic source use and Paleoarchaic foraging

territories in the Great Basin. American Antiquity 68: 5–38.

Jones, T. L., Fitzgerald, R. T., Kennett, D. J., Miksicek, C. H., Fagan, J. L., Sharp, J., and Erlandson, J. M.

(2002). The Cross Creek site (CA-SLO-1797) and its implications for New World colonization.

American Antiquity 67: 213–230.

Joyce, D. J. (2006). Chronology and new research on the Schaefer mammoth (?Mammuthus primigenius)site, Kenosha County, Wisconsin, USA. Quaternary International 142–143: 44–57.

Joyce, D., and Blazina-Joyce, R. (2002). A chronological assessment of the Schaefer mammoth site,

southeastern Wisconsin. Current Research in the Pleistocene 19: 43–45.

Keefer, D. K., deFrance, S. D., Moseley, M. E., Richardson III, J. B., Satterlee, D. R., and Day-Lewis, A.

(1998). Early maritime economy and El Nino events at Quebrada Tacahuay, Peru. Science 281:1833–1835.

Kelly, R. L. (1995). The Foraging Spectrum: Diversity in Hunter-Gatherer Lifeways, Smithsonian

Institution Press, Washington, DC.

Kelly, R. L. (2003). Maybe we do know when people first came to North America; and what does it mean

if we do? Quaternary International 109–110: 133–145.

Kemp, B. M., Malhi, R. S., McDonough, J., Bolnick, D. A., Eshleman, J. A., Rickards, O., Martinez-

Labarga, C., Johnson, J. R., Lornez, J. G., Dixon, E. J., Fifield, T. E., Heaton, T. H., Worl, R., and

Smith, D. G. (2007). Genetic analysis of early Holocene skeletal remains from Alaska and its

implications for the settlement of the Americas. American Journal of Physical Anthropology 132:1–17.

366 J Archaeol Res (2011) 19:327–375

123

Page 41: fulltext (2)

Kitchen, A., Miyamoto, M. M., and Mulligan, C. J. (2008). A three-stage colonization model for the

peopling of the Americas. PLoS 3: 1–7.

Kooyman, B., Hills, L. V., McNeil, P., and Tolman, S. (2006). Late Pleistocene horse hunting at the

Wally’s Beach site (DhPg-8), Canada. American Antiquity 71: 101–121.

Kuzmin, Y. V., and Keates, S. G. (2005). Dates are not just data: Paleolithic settlement patterns in Siberia

derived from radiocarbon records. American Antiquity 70: 773–789.

Kuzmin, Y. V., Kosintsev, P. A., Razhev, D. I., and Hodgins, G. W. L. (2009). The oldest directly-dated

human remains in Siberia: AMS 14C age of talus bone from the Baigara locality, West Siberian

Plain. Journal of Human Evolution 57: 91–95.

Lanata, J. L., Martino, L., Osella, A., and Garcia-Herbst, A. (2008). Demographic conditions necessary to

colonize new spaces: The case for early human dispersal in the Americas. World Archaeology 40:520–537.

Largent, F., Jr. (2004). Early Americans in eastern Beringia: Pre-Clovis traces at Swan Point, Alaska.

Mammoth Trumpet 20: 4–7.

Lavalle, D., and Bahn, P. G. (2000). The First South Americans: The Peopling of a Continent from theEarliest Evidence to High Culture, University of Utah Press, Salt Lake City.

Lell, J. T., Sukernik, R. I., Starikovskaya, Y. B., Su, B., Jin, L., Schurr, T. G., Underhill, A. P., and

Wallace, D. C. (2002). The dual origin and Siberian affinities of Native American Y chromosomes.

American Journal of Human Genetics 70: 192–206.

Leonard, J. A., Wayne, R. K., and Cooper, A. (2000). Population genetics of Ice Age brown bears.

Proceedings of the National Academy of Sciences 97: 1651–1654.

Lewis, R. B. (2000). Sea-level rise and subsidence effects on Gulf Coast archaeological site distributions.

American Antiquity 65: 525–541.

Macaulay, V., Hill, C., Achilli, A., Rengo, C., Clarke, D., Meehan, W., Blackburn, J., Semino, O.,

Scozzari, R., Cruciani, F., Taha, A., Shaari, N. K., Raja, J. M., Ismail, P., Zainuddin, Z., Goodwin,

W., Bulbeck, D., Bandelt, H., Oppenheimer, S., Torroni, A., and Richards, M. (2005). Single, rapid

coastal settlement of Asia revealed by analysis of complete mitochondrial genomes. Science 308:1034–1036.

MacDonald, D. H. (2004). Early Paleoindians as estate settlers: Archaeological, ethnographic, and

evolutionary insights into the peopling of the New World. In Barton, C. M., Clark, G. A., Yesner, D.

R., and Pearson, G. A. (eds.), The Settlement of the American Continents: A MultidisciplinaryApproach to Human Biogeography, University of Arizona Press, Tucson, pp. 173–176.

Macphail, R. I., and McAvoy, J. M. (2008). A micromorphological analysis of stratigraphic integrity and

site formation at Cactus Hill, and Early Paleoindian and hypothesized pre-Clovis occupation in

south-central Virginia, USA. Geoarchaeology 23: 675–694.

Malhi, R. S., Eshleman, J. A., Greenberg, J. A., Weiss, D. A., Shook, B. A. S., Kaestle, F. A., Lorenz, J.

G., Kemp, B. M., Johnson, J. R., and Smith, D. G. (2002). The structure of diversity within New

World mitochondrial DNA haplogroups: Implications for the prehistory of North America.

American Journal of Human Genetics 70: 905–919.

Malhi, R. S., Kemp, B. M., Eshleman, J. A., Cybulski, J., Smith, D. G., Cousins, S., and Harry, H. (2007).

Mitochondrial haplogroup M discovered in prehistoric North Americans. Journal of ArchaeologicalScience 34: 642–648.

Malhi, R. S., and Smith, D. G. (2002). Brief communication: Haplogroup X confirmed in prehistoric

North America. American Journal of Physical Anthropology 119: 84–86.

Mandryk, C. A. S., Josenhans, H., Fedje, D. W., and Matherwes, R. W. (2001). Late Quaternary

paleoenvironments of northwest North America: Implications for inland versus coastal migration

routes. Quaternary Science Reviews 20: 301–314.

Marshall, E. (2001). Pre-Clovis sites fight for acceptance. Science 291: 1730–1733.

McAvoy, J. M., and McAvoy, L. D. (eds.) (1997). Archaeological Investigations of Site 44SX202, CactusHill, Sussex County, Virginia, Research Report Series No. 8, Virginia Department of Historic

Resources, Richmond.

Meltzer, D. J. (2003). Lessons in landscape learning. In Rockman, M., and Steele, J. (eds.), Colonizationof Unfamiliar Landscapes: The Archaeology of Adaptation, Routledge, New York, pp. 222–241.

Meltzer, D. (2004). Peopling of North America. In Gillespie, A. R., Porter, S. C., and Atwater, B. F.

(eds.), The Quaternary Period in the United States: Developments in Quaternary Science, Elsevier,

San Diego, CA, pp. 539–563.

Meltzer, D. J. (2009). First Peoples in a New World, University of California Press, Berkeley.

J Archaeol Res (2011) 19:327–375 367

123

Page 42: fulltext (2)

Meltzer, D. J., Grayson, D. K., Ardila, G., Barker, A. W., Dincauze, D. F., Haynes, Jr., C. V., Mena, F.,

Nunez, L., and Stanford, D. J. (1997). On the Pleistocene antiquity of Monte Verde, southern Chile.

American Antiquity 62: 659–663.

Mena, F., Reyes, O., Stafford, Jr., T. W., and Southon, J. (2003). Early human remains from Bano Nuevo-

1 cave (central Patagonian Andes, Chile). Quaternary International 109–110: 113–121.

Merriwether, D. A. (2006). Mitochondrial DNA. In Ubelaker, D. H. (ed.), Handbook of North AmericanIndians, V. 3, Sturtevant, W. C., and Stanford, D. (eds.), Environment, Origins, and Population,

Smithsonian Institution Press, Washington, DC, pp. 817–830.

Miotti, L. L. (2003). Patagonia: A paradox for building images of the first Americans during Pleistocene/

Holocene transition. Quaternary International 109–110: 147–173.

Miotti, L., and Salemme, M. C. (2003). When Patagonia was colonized: People mobility at high latitudes

during Pleistocene/Holocene transition. Quaternary International 109–110: 95–111.

Miotti, L., Salemme, M., and Rabass, J. (2002). Radiocarbon chronology at Piedra Museo locality. In

Miotti, L., Salemme, M., and Flegenheimer, N. (eds.), Ancient Evidence for Paleo South Americans:From Where the South Winds Blow, Center for the Study of the First Americans, Texas A&M

University, College Station, pp. 99–104.

Mishmar, D., Ruiz-Pesini, E., Golik, P., Macaulay, V., Clark, A. G., Hosseini, S., Brandon, M., Easley,

K., Chen, E., Brown, M. D., Sukernik, R. I., Olckers, A., and Wallace, D. C. (2003). Natural

selection shaped regional mtDNA variation in humans. Proceedings of the National Academy ofSciences 100: 171–176.

Mochanov, Y. A., and Fedoseeva, S. A. (1996). Aldanks: Aldan River Valley, Sakha Republic. In West,

F. H. (ed.), American Beginnings: The Prehistory and Palaeoecology of Beringia, University of

Chicago Press, Chicago, pp. 157–214.

Montenegro, A., Hetherington, R., Eby, M., and Weaver, A. J. (2006). Modeling pre-historic transoceanic

crossings into the Americas. Quaternary Science Reviews 25: 1323–1338.

Moore, J. H., and Moseley, M. E. (2001). How many frogs does it take to leap around the Americas?

Comments on Anderson and Gillam. American Antiquity 66: 526–529.

Morlan, R. E. (2003). Current perspectives on the Pleistocene archeology of eastern Beringia. QuaternaryResearch 60: 123–132.

Neves, W. A., and Blum, M. (2000). The Buhl Burial: A comment on Green et al. American Antiquity 65:191–193.

Neves, W. A., and Hubbe, M. (2005). Cranial morphology of early Americans from Lagoa Santa, Brazil:

Implications for the settlement of the New World. Proceedings of the National Academy of Sciences102: 18309–18314.

Neves, W. A., Prous, A., Gonzalez-Jose, R., Kipnis, R., and Powell, J. (2003). Early Holocene human

skeletal remains from Santana do Riacho, Brazil: Implications for the settlement of the New World.

Journal of Human Evolution 45: 19–42.

Neves, W. A., Gonzalez-Jose, R., Hubbe, M., Kipnis, R., Araujo, A. G. M., and Blasi, O. (2004). Early

Holocene human skeletal remains from Cerca Grande, Lagoa Santa, central Brazil, and the origins

of the first Americans. World Archaeology 36: 479–501.

Neves, W. A., Hubbe, M., and Pilo, L. B. (2007). Early Holocene human skeletal remains from

Sumidouro Cave, Lagoa Santa, Brazil: History of discoveries, geological and chronological context,

and comparative cranial morphology. Journal of Human Evolution 52: 16–30.

Nichols, J. (2008). Language spread rates and prehistoric American migration rates. CurrentAnthropology 49: 1109–1118.

O’Rourke, D. (2009). Human migrations: The two roads taken. Current Biology 19: R203–R205.

O’Rourke, D. H., Hayes, M. G., and Carlyle, S. W. (2000) Ancient DNA studies in physical anthropology.

Annual Review of Anthropology 29: 217–242.

Overstreet, D. F. (2005). Late-glacial ice-marginal adaptation in southeastern Wisconsin. In Bonnichsen,

R., Lepper, B. T., Stanford, D., and Waters, M. R. (eds.), Paleoamerican Origins: Beyond Clovis,

Center for the Study of the First Americans, Texas A&M University, College Station, pp. 183–195.

Overstreet, D., and Kolb, M. (2003). Geoarchaeological contexts for Late Pleistocene archaeological sites

with human-modified wooly mammoth remains in southeastern Wisconsin, USA. Geoarchaeology18: 91–114.

Owsley, W. D., and Jantz, R. L. (2001). Archaeological politics and public interest in Paleoamerican

studies: Lessons from Gordon Creek Woman and Kennewick Man. American Antiquity 66:565–575.

368 J Archaeol Res (2011) 19:327–375

123

Page 43: fulltext (2)

Pavlov, P., Svendsen, J. I., and Indrelid, S. (2001). Human presence in the European Arctic nearly

40,000 years ago. Nature 413: 64–67.

Perego, U. A., Achilli, A., Angerhofer, N., Accetturo, M., Pala, M., Olivieri, A., Kashani, B. H., Ritchie,

K. H., Scozzari, R., Kong, Q.-P., Myres, N. M., Salas, A., Semino, O., Bandelt, Hans-Jurgen,

Woodward, S. R., and Torroni, A. (2009). Distinctive Paleo-Indian migration routes from Beringia

marked by two rare mtDNA haplogroups. Current Biology 19: 1–8.

Peros, M. C., Munoz, S. E., Gajewski, K., and Viau, A. E. (2010). Prehistoric demography of North

America inferred from radiocarbon data. Journal of Archaeological Science 37: 656–664.

Pitulko, V. V., Nikolsky, P. A., Girya, E. Y., Basilyan, A. E., Tumskoy, V. E., Koulakov, S. A., Astakhov,

S. N., Pavlova, E. Y., and Anisimov, M. A. (2004). The Yana RHS site: Humans in the Arctic before

the last glacial maximum. Science 303: 52–56.

Poinar, H., Fiedel, S., King, C. E., Devault, A. M., Bos, K., Kuch, M., and Debruyne, R. (2009). Comment

on ‘‘DNA from Pre-Clovis human coprolites in Oregon, North America.’’ Science 325: 148a.

Powell, J. F. (2005). The First Americans: Race, Evolution, and the Origin of Native Americans,

Cambridge University Press, Cambridge.

Ramsey, C. L., Griffiths, P. A., Fedge, D. W., Wigen, R. J., and Mackie, Q. (2004). Preliminary

investigation of a late Wisconsinan fauna from K1 cave, Queen Charlotte Islands (Haida Gwaii),

Canada. Quaternary Research 62: 105–109.

Rasmussen, M., Cummings, L. S., Gilbert, M. T. P., Bryant, V., Smith, C., Jenkins, D. J., and Willerslev,

E. (2009). Response to Comment by Goldberg et al. on ‘‘DNA from Pre-Clovis human coprolites in

Oregon, North America.’’ Science 325: 148d.

Ray, N., and Excoffier, L. (2009). Inferring past demography using spatially explicit population genetic

models. Human Biology 81: 141–157.

Redmond, B. G., and Tankersley, K. B. (2005). Evidence of early Paleoindian bone modification and use

at the Sheriden Cave site (33WY252), Wyandot County, Ohio. American Antiquity 70: 503–526.

Reidla, M., Kivisild, T., Metspalu, E., Kaldma, K., Tambets, K., Tolk, H., Parik, J., Loogvali, E.,

Derenko, M., Malyarchuk, B., Bermisheva, M., Zhadanov, S., Pennarun, E., Gubina, M.,

Golubenko, M., Damba, L., Federova, S., Gusar, V., Grechanina, E., Mikerezi, I., Moisan, J.,

Chaventre, A., Khusnutdinova, E., Osipova, L., Stepanov, V., Voevoda, M., Achilli, A., Rengo, C.,

Rickards, O., De Stefano, G. F., Papiha, S., Beckman, L., Jaincijevic, B., Rudan, P., Anagnou, N.,

Michalodimitrakis, E., Koziel, S., Usanga, E., Geberhiwot, T., Herrnstadt, C., Howell, N., Torroni,

A., and Villiams, R. (2003). Origin and diffusion of mtDNA haplogroup X. The American Society ofHuman Genetics 73: 1178–1190.

Richards, M. P., Pettitt, P. B., Stiner, M. C., and Trinkaus, E. (2001). Stable isotope evidence for

increasing dietary breadth in the European mid-Upper Paleolithic. Proceedings of the NationalAcademy of Sciences 98: 6528–6532.

Rick, T. C., Erlandson, J. M., and Vellanoweth, R. L. (2001). Paleocoastal marine fishing on the Pacific

Coast of the Americas: Perspectives from Daisy Cave, California. American Antiquity 66: 595–613.

Roper, D. C., and Wygal, B. T. (2002). The spatial component of the western Clovis chronology. CurrentResearch in the Pleistocene 19: 76–78.

Sandweiss, D. H., Richardson III, J. B., Reitz, E. J., Hsu, J. T., and Feldman, R. A. (1989). Early marine

adaptations in the Andes: Preliminary studies at the Ring site, Peru. Annual Review of Anthropology32: 239–262.

Sandweiss, D. H., McInnis, H., Burger, R. L., Cano, A., Ojeda, B., Paredes, R., del Carmen Sandweiss,

M., and Glascock, M. D. (1998). Quebrada Jaguay: Early South American maritime adaptations.

Science 281: 1830–1832.

Santos, F. R., Pandya, A., Tyler-Smith, C., Pena, S. J. D., Schanfield, M., Leonard, W. R., Osipova, L.,

Crawford, M. H., and Mitchell, R. J. (1999). The central Siberian origin for Native American Y

chromosomes. American Journal of Human Genetics 64: 619–628.

Sardi, M. L., Ramırez Rozzi, F., Gonzalez-Jose, R., and Pucciarelli, H. M. (2005). South Amerindian

craniofacial morphology: Diversity and implications for Amerindian evolution. American Journal ofPhysical Anthropology 128: 747–756.

Scheinsohn, V. (2003). Hunter-gatherer archaeology in South America. Annual Review of Anthropology32: 339–361.

Schroeder, K. B., Schurr, T. G., Long, J. C., Rosenburg, N. A., Crawford, M. H., Tarskaia, L. A., Osipova,

L. P., Zhadanov, S. I., and Smith, D. G. (2007). A private allele ubiquitous in the Americas. BiologyLetters 3: 218–223.

J Archaeol Res (2011) 19:327–375 369

123

Page 44: fulltext (2)

Schroeder, K. B., Jakobsson, M., Crawford, M. H., Schurr, T. G., Boca, S. M., Conrad, D. F., Tito, R. Y.,

Osipova, L. P., Tarskaia, L. A., Zhadanov, S. I., Wall, J. D., Pritchard, J. K., Malhi, R. S., Smith, D.

G., and Rosenberg, N. A. (2009). Haplotypic background of a private allele at high frequency in the

Americas. Molecular Biology and Evolution 26: 995–1016.

Schurr, T. G. (2004). The peopling of the New World: Perspectives from molecular anthropology. AnnualReview of Anthropology 33: 551–583.

Schurr, T. G., and Sherry, S. T. (2004). Mitochondrial DNA and Y chromosome diversity and the

peopling of the Americas: Evolutionary and demographic evidence. American Journal of HumanBiology 16: 420–439.

Seielstad, M., Yuldasheva, N., Singh, N., Underhill, P., Oefner, P., Shen, P., and Wells, S. (2003) A novel

Y-chromosome variant puts an upper limit on the timing of the first entry into the Americas.

American Journal of Human Genetics 73: 700–705.

Shapiro, B., Drummond, A. J., Rambaut, A., Wilson, M. C., Matheus, P. E., Sher, A. V., Pybus, O. G.,

Gilbert, M. T. P., Barnes, I., Binladen, J., Willerslev, E., Hansen, A. J., Baryshnikov, G. F., Burns, J.

A., Davydov, S., Driver, J. C., Froese, D. G., Harington, C. R., Keddie, G., Kosintsev, P., Kunz, M.

L., Martin, L. D., Stephenson, R. O., Storer, J., Tedford, R., Zimov, S., and Cooper. A. (2004). Rise

and fall of the Beringian steppe bison. Science 306: 1561–1565.

Silva, Jr., W. A., Bonatto, S. L., Holanda, A. J., Riberio-dos Santos, A. K., Paixao, B. M., Goldman, G.

H., Abe-Sandes, K., Rodriguez-Delfin, L., Barbosa, M., Paco-Larson, M. L., Petzl-Erler, M. L.,

Valente, V., Santos, S. E. B., and Zago, M. A. (2002). Mitochondrial genome diversity of Native

Americans supports a single early entry of founder populations into America. American Journal ofHuman Genetics 71: 187–192.

Starikovskaya, E. B., Sukernik, R. I., Derbenva, O. A., Volodko, N. V., Ruiz-Pesini, E., Torroni, A.,

Brown, M. D., Lott, M. T., Hosseini, S. H., Huoponen, K., and Wallace, D. C. (2005). Mitochondrial

DNA diversity in indigenous populations of the southern extent of Siberia, and the origins of Native

American haplogroups. Annals of Human Genetics 69: 67–89.

Steele, D. G., and Powell, J. F. (2002). Facing the past: A view of the North American human fossil

record. In Jablonski, N. G. (ed.), The First Americans: The Pleistocene Peopling of the New World,

Memoirs No. 27, California Academy of Sciences, San Francisco, pp. 93–122.

Steele, J. (2009). Human dispersals: Mathematical models and the archaeological record. Human Biology81: 121–140.

Steele, J., and Politis, G. (2009). AMS 14C dating of early human occupation of southern South America.

Journal of Archaeological Science 36: 419–429.

Steele, J., Adams, J., and Sluckin, T. (1998). Modelling Paleoindian dispersals. World Archaeology 30:286–305.

Steffy, K., and Goodyear, A. C. (2006). Clovis macro blades from the Topper site, 38AL23, Allendale

County, South Carolina. Current Research in the Pleistocene 23: 147–149.

Straus, L. G. (2000). Solutrean settlement of North America? A review of reality. American Antiquity 65:219–226.

Straus, L. G., Bicho, N., and Winegardner, A. C. (2000). The Upper Paleolithic settlement of Iberia: First-

generation maps. Antiquity 74: 553–566.

Straus, L. G., Meltzer, D. J., and Goebel, T. (2005). Ice Age Atlantis? Exploring the Solutrean-Clovis

‘‘connection.’’ World Archaeology 37: 507–532.

Surovell, T. A. (2000). Early Paleoindian women, children, mobility, and fertility. American Antiquity 65:493–508.

Surovell, T. A. (2003). Stimulating coastal migration in New World colonization. Current Anthropology44: 580–591.

Surovell, T. A., and Waguespack, N. M. (2008). How many elephant kills are 14? Clovis mammoth and

mastodon kills in context. Quaternary International 191: 82–97.

Surovell, T. A., and Waguespack, N. M. (2009). Human prey choice in the Late Pleistocene and its

relation to megafaunal extinctions. In Haynes, G. (ed.), American Megafaunal Extinctions at the Endof the Pleistocene, Springer, New York, pp. 77–105.

Swedlund, A., and Anderson, D. (2003). Gordon Creek Woman meets Spirit Cave Man: A response to

Comment by Owsley and Jantz. American Antiquity 68: 161–167.

Tamm, E., Kivisild, T., Reidla, M., Metspalu, M., Smith, D. G., Mulligan, C. J., Bravi, C. M., Rickards,

O., Martinez-Labarga, C., Khusnutdinova, E. K., Federova, S. A., Golubenko, M. V., Sepanov, V.

A., Gubina, M. A., Zhadanov, S. I., Ossipova, L. P., Damba, L., Voevoda, M. I., Dipierri, J. E.,

370 J Archaeol Res (2011) 19:327–375

123

Page 45: fulltext (2)

Villems. R., and Malhi, R. S. (2007). Beringian standstill and spread of Native American founders.

PLoS ONE September 2007: e829.

Tarazona-Santos, E., and Santos, F. R. (2002). The peopling of the Americas: A second major migration?

American Journal of Human Genetics 70: 1377–1380.

Torroni, A., Schurr, T. G., Yang, C.-C., Szathmary, E. J. E., Williams, R. C., Schanfield, M. S., Troup, G.

A., Knowler, W. C., Lawrence, D. N., Weiss, K. M., and Wallace, D. C. (1992). Native American

mitochondrial DNA analysis indicates that the Amerind and the Na-Dene populations were founded

by two independent migrations. Genetics 130: 153–162.

Torroni, A., Schurr, T. G., Cabell, M. F., Brown, M. D., Neel, J. V., Larsen, M., Smith, D. G., Vullo, C.

M., and Wallace, D. C. (1993). Asian affinities and the continental radiation of the four founding

Native American mtDNAs. American Journal of Human Genetics 53: 563–590.

Turner II, C. G. (2002). Teeth, needles, dogs and Siberia. In Jablonski, N. G. (ed.), The First Americans:The Pleistocene Peopling of the New World, Memoirs No. 27, California Academy of Sciences, San

Francisco, pp. 123–158.

Turner II, C. G. (2003). Three ounces of sea shells and one fish bone do not a coastal migration make.

American Antiquity 68: 391–395.

Vallanoweth, R. L., Lambright, M. R., Erlandson, J. M., and Rick, T. C. (2003). Early New World

maritime technologies: Sea grass cordage, shell beads, and a bone tool from Cave of the Chimneys,

San Miguel Island, California, USA. Journal of Archaeological Science 30: 1161–1173.

Vereschagin, N. K., and Ukraintseva, V. V. (1985). Proiskhozhdenie I stratigrafiya Berelekhskogo

‘‘kladbishcha’’ mamontov. Trudy Zoologicheskogo Institute AN SSSR 131: 104–113.

Wagner, D. P., and McAvoy, J. M. (2004). Pedoarchaeology of Cactus Hill, a sandy Paleoindian site in

Southeastern Virginia, U.S.A. Geoarchaeology 19: 297–322.

Waguespack, N. M. (2005). The organization of male and female labor in foraging societies: Implications

for early Paleoindian archaeology. American Anthropologist 107: 666–676.

Waguespack, N. M. (2007). Why we’re still arguing about the Pleistocene occupation of the Americas.

Evolutionary Anthropology 16: 63–74.

Waguespack, N. M., and Surovell, T. A. (2003). Clovis hunting strategies, or how to make out on

plentiful resources. American Antiquity 68: 333–352.

Wang, S., Lewis, Jr., C. M., Jakobsson, M., Ramachandran, S., Ray, N., Bedoya, G., Rojas, W., Parra, M.

V., Molina, J. A., Gallo, C., Mazzotti, G., Poletti, G., Hill, K., Hurtado, A. M., Labuda, D., Klitz,

W., Barrantes, R., Bortolini, M. C., Salazano, F. M., Petzl-Erler, M. L., Tsuneto, L. T., Llop, E.,

Rothhammer, F., Excoffier, L., Feldman, M. W., Rosenberg, N. A., and Ruiz-Linares, A. (2007).

Genetic variation and population structure in Native Americans. PLoS 3: 2049–2067.

Waters, M. R., and Stafford, Jr., T. W. (2007a). Redefining the Age of Clovis: Implications for the

peopling of the Americas. Science 315: 1122–1126.

Waters, M. R., and Stafford, Jr., T. W. (2007b). Response to comment on ‘‘Redefining the age of Clovis:

Implications for the peopling of the Americas.’’ Science 317: 320c.

Waters, M. R., Stafford, Jr., T. W., Redmond, B. G., and Tankersley, K. B. (2009a). The age of the

Paleoindian assemblage at Sheriden Cave, Ohio. American Antiquity 74: 107–111.

Waters, M. R., Forman, S. L., Stafford, Jr., T. W., and Foss, J. (2009b). Geoarchaeological investigations

at the Topper and Big Pine Tree sites, Allendale County, South Carolina. Journal of ArchaeologicalScience 36: 1300–1311.

Wells, R. S., Yuldasheva, N., Ruzibakiev, R., Underhill, P. A., Evseeva, I., Blue-Smith, J., Jin, L., Su, B.,

Pitchappan, R., Shanmugalakshmi, S., Balakrishnan, K., Read, M., Pearson, N. M., Zerjal, T.,

Webster, M. T., Zholoshvili, I., Jamarjashvili, E., Gombarov, S., Nikbin, B., Dostiev, A.,

Aknazarov, O., Zalloua, P., Tsoy, I., Kitaev, M., Mirrakhimov, M., Chariev, A., and Bodmer, W. F.

(2001). The Eurasian heartland: A continental perspective on Y-chromosome diversity. Proceedingsof the National Academy of Sciences 98: 10244–10249.

Willerselv, E., Hansen, A. J., Binladen, J., Brand, T. B., Gilbert, M. T. P., Shapiro, B., Bunce, M., Wiuf,

C., Gilichinsky, D. A., and Cooper, A. (2003). Diverse plant and animal genetic records from

Holocene and Pleistocene sediments. Science 300: 791–795.

Wilson, M. C. (1996). Late Quaternary vertebrates and the opening of the ice-free corridor, with special

reference to the genus Bison. Quaternary International 32: 97–105.

Wobst, H. M. (1974). Boundary conditions for Paleolithic social systems: A simulation approach.

American Antiquity 39: 147–178.

Wobst, H. M. (1976). Locational relationships in Paleolithic societies. Journal of Human Evolution 5:49–58.

J Archaeol Res (2011) 19:327–375 371

123

Page 46: fulltext (2)

Wyatt, S. (2004). Ancient transpacific voyaging to the New World via Pleistocene South Pacific Islands.

Geoarchaeology 19: 511–529.

Yesner, D. R. (1980). Maritime hunter-gatherers: Ecology and prehistory. Current Anthropology 21:727–750.

Yesner, D. R. (2001). Human dispersal into interior Alaska: Antecedent conditions, mode of colonization,

and adaptations. Quaternary Science Reviews 20: 315–327.

Yesner, D. R., Barton, M. C., Clark, G. A., and Pearson, G. A. (2004). Peopling of the Americas and

continental colonization: A millennial perspective. In Barton, C. M., Clark, G. A., Yesner, D. R., and

Pearson, G. A. (eds.), The Settlement of the American Continents: A Multidisciplinary Approach toHuman Biogeography, University of Arizona Press, Tucson, pp. 196–213.

Zazula, G. D., Froese, D. G., Schweger, C. E., Mathewes, R. W., Beaudoin, A. B., Telka, A. M.,

Harington, C. R., and Westgate, J. A. (2003). Ice-age steppe vegetation in east Beringia. Nature 426:603.

Zegura, S. L., Karafet, T. M., Zhivotovsky, L. A., and Hammer, M. F. (2004). High resolution SNPs and

microsatellite haplotypes point to a single, recent entry of Native American Y-chromosomes into the

Americas. Molecular Biology and Evolution 21: 164–175.

Bibliography of recent literature

Alroy, J. (2001). A multispecies overkill simulation of the End-Pleistocene megafaunal mass extinction.

Science 292: 1893–1896.

Alroy, J. (2001). Response to Grayson, D. K., ‘‘Did human hunting cause mass extinction?’’ Science 294:1459–1460.

Alroy, J. (2001). Response to Slaughter, R. and Skulan, J., ‘‘Did human hunting cause mass extinction?’’

Science 294: 1461–1462.

Ballard, R. D. (2008). Archaeological Oceanography, Princeton University Press, Princeton, NJ.

Barnosky, A. D. (2004). Effect of climate change on terrestrial vertebrate biodiversity. In Barnosky, A. D.

(ed.), Biodiversity Responses to Climate Change in the Middle Pleistocene: The Porcupine CaveFauna from Colorado, University of California Press, Berkeley, pp. 341–371.

Barnosky, A. D., Koch, P. L., Feranec, R. S., Wing, S. L., and Shabel, A. B. (2004). Assessing the causes

of Late Pleistocene extinctions on the continents. Science 306: 70–75.

Barton, C. M., Clark, G. A., Yesner, D. R., and Pearson, G. A. (eds.) (2004). The Settlement of theAmerican Continents: A Multidisciplinary Approach to Human Biogeography, University of

Arizona Press, Tucson.

Beck, C., and Jones, G. T. (2008). The Archaeology of the Eastern Nevada Paleoarchaic, Part 1: TheSunshine Locality, University of Utah Press, Salt Lake City.

Boeskorov, G. G. (2006). Arctic Siberia: Refuge of the mammoth fauna in the Holocene. QuaternaryInternational 142: 119–123.

Bonnichsen, R., and Turnmire, K. L. (eds.) (2005). Ice Age Peoples of North America: Environments,Origins and Adaptations, Texas A&M University Press, College Station.

Bonnichsen, R., Lepper, B. T., Stanford, D., and Waters, M. R. (eds.). (2006). Paleoamerican Origins:Beyond Clovis, Texas A&M University Press, College Station.

Bourque, B. J., Cox, S. L., and Whitehead, R. H. (2001). Twelve Thousand Years: American Indians inMaine, University of Nebraska Press, Lincoln.

Brunswig, R. H., and Pitblado, B. L. (eds.) (2007). Frontiers in Paleoindian Archaeology: From the DentSite to the Rocky Mountains, University Press of Colorado, Niwot.

Buchanan, B., Collard, M., and Edinborough, K. (2008). Paleoindian demography and the extraterrestrial

impact hypothesis. Proceedings of the National Academy of Sciences 105: 11651–11654.

Burney, D. A., and Flannery, T. F. (2005). Fifty millennia of catastrophic extinctions after human contact.

Trends in Ecology and Evolution 20: 395–401.

Chatters, J. C. (2002). Ancient Encounters: Kennewick Man and the First Americans, Simon and

Schuster, New York City.

Collins, M. B. (2002). Clovis Blade Technology: A Comparative Study of the Keven Davis Cache, Texas,

University of Texas Press, Austin.

Dillehay, T. D. (2000). The Settlement of the Americas: A New Prehistory, Basic Books, New York.

372 J Archaeol Res (2011) 19:327–375

123

Page 47: fulltext (2)

Downey, R. (2000). Riddle of the Bones: Politics, Science, Race, and the Story of Kennewick Man,

Springer, New York.

Ellis, C., and Deller, D. B. (2001). An Early Paleo-Indian Site Near Parkhill, Ontario, University of

Washington Press, Seattle.

Fiedel, S., and Haynes, G. (2004). A premature burial: Comments on Grayson and Meltzer’s ‘‘Requiem

for overkill.’’ Journal of Archaeological Science 31: 121–131.

Gillespie, J., Tupakka, S., and de Mille, C. (eds.) (2001). On Being First: Cultural Innovation andEvironmental Consequences of First Peopling, Proceedings of the 31st Annual Chacmool

Conference, Archaeological Association of the University of Calgary, Calgary, Alberta.

Graf, K. E., and Schmitt, D. N. (eds.) (2007). Paleoindian or Paleoarchaic? Great Basin Human Ecologyat the Pleistocene – Holocene Transition, University of Utah Press, Salt Lake City.

Grayson, D. K. (2001). Did human hunting cause mass extinction? Science 294: 1459.

Grayson, D. K. (2001). The archaeological record of human impacts on animal populations. Journal ofWorld Prehistory 15: 1–68.

Grayson, D. K., and Meltzer, D. J. (2002). Clovis hunting and large mammal extinction: A critical review

of the evidence. Journal of World Prehistory 16: 313–359.

Grayson, D. K., and Meltzer, D. J. (2003). A requiem for North American overkill. Journal ofArchaeological Science 30: 585–593.

Grayson, D. K., and Meltzer, D. J. (2004). North American overkill continued? Journal of ArchaeologicalScience 31: 133–136.

Guthrie, R. D. (2006). New carbon dates link climatic change with human colonization and Pleistocene

extinctions. Nature 441: 207–209.

Haynes, Jr., C. V. (2008). Younger Dryas ‘‘black mats’’ and the Rancholabrean termination in North

America. Proceedings of the National Academy of Sciences 105: 6520–6525.

Haynes, C. V., and Huckell, B. B. (eds.) (2007). Murray Springs: A Clovis Site with Multiple ActivityAreas in the San Pedro Valley, Arizona, University of Arizona Press, Tucson.

Haynes, G. (2002). The catastrophic extinction of North American mammoths and mastodonts. WorldArchaeology 33: 391–416.

Haynes, G. (2009). American Megafaunal Extinctions at the End of the Pleistocene, Springer, New York.

Holdaway, R. N., and Jacomb, C. (2000). Rapid extinction of the Moas (Aves: Dinornithiformes): Model,

test, and implications. Science, New Series 287: 2250–2254.

Holliday, V. T. (2000). The evolution of Paleoindian geochronology and typology on the Great Plains.

Geoarchaeology 15: 227–290.

Holliday, V. T. (2003). Where have all the mammoth gone? Science 300: 1373–1374.

Huckell, B. B., and Kilby, J. D. (eds.) (2004). Readings in Late Pleistocene North America and EarlyPaleoindians: Selections from American Antiquity, Society for American Archaeology, Washington,

DC.

Jablonski, N. G. (ed.) (2002). The First Americans: The Pleistocene Colonization of the New World,

University of California Press, Berkeley.

Johnson, C. N. (2005). What can the data on late survival of Australian megafauna tell us about the cause

of their extinction? Quaternary Science Review 24: 2167–2172.

Kennett, D. J., Kennett, J. P., West, A., Mercer, C., Que Hee, S. S., Bement, L., Bunch, T. E., Sellers, M.,

and Wolbach, W. S. (2009). Nanodiamonds in the Younger Dryas boundary sediment layer. Science323: 94.

Kooyman, B., Newman, M. E., Cluney, C., Lobb, M., Tolman, S., McNeil, P., and Hills, L. V. (2001).

Identification of horse exploitation by Clovis hunters based on protein analysis. American Antiquity66: 686–691.

Kooyman, B., Hills, L. V., McNeil, P., and Tolman, S. (2006). Late Pleistocene horse hunting at the

Wally’s Beach site (DhPg-8), Canada. American Antiquity 71: 101–121.

Kunz, M. L. (2003). The Mesa Site: Paleoindians above the Arctic Circle, US Department of the Interior,

Bureau of Land Management, Alaska State Office, Anchorage.

Larson, M. L., Kornfeld, M., and Frison, G. C. (eds.) (2009). Hell Gap: A Stratified Paleoindian Campsiteat the Edge of the Rockies, University of Utah Press, Salt Lake City.

Lepper, B. T., and Bonnichsen, R. (eds.) (2004). New Perspectives on the First Americans, Texas A&M

University Press, College Station.

MacNeish, R. S., and Libby, J. G. (eds.) (2003). Pendejo Cave, University of New Mexico Press,

Albuquerque.

J Archaeol Res (2011) 19:327–375 373

123

Page 48: fulltext (2)

Madsen, D. B. (ed.) (2004). Entering America: Northeast Asia and Beringia before the Last GlacialMaximum, University of Utah Press, Salt Lake City.

Marlon, J. R., Bartlein, P. J., Walsh, M. K., Harrison, S. P., Brown, K. J., Edwards, M. E., Higuera, P. E.,

Power, M. J., Anderson, R. S., Briles, C., Brunelle, A., Carcaillet, C., Daniels, M., Hu, F. S., Lavois,

M., Long, C., Minckley, T., Richard, P. J. H., Scott, A. C., Shafer, D. S., Tinner, W., Umbanhowar,

Jr., C. E., and Whitlock, C. (2009). Wildfire responses to abrupt climate change in North America.

Proceedings of the National Academy of Sciences 106: 2519–2524.

Martin, P. S. (2005). Twilight of the Mammoths: Ice Age Extinctions and the Rewilding of America,

University of California Press, Berkeley.

McNeil, P., Hills, L. V., Kooyman, B., and Toman, S. M. (2005). Mammoth tracks indicate a declining

Late Pleistocene population in southwestern Alberta, Canada. Quaternary Science Reviews 24:1253–1259.

Moriarty, K. C., McCulloch, M. T., Wells, R. T., and McDowell, M. C. (2000). Mid-Pleistocene cave fills,

megafaunal remains and climate change at Naracoorte, South Australia: Towards a predictive model

using U-Th dating of speleothems. Palaeogeography, Palaeoclimatology, Palaeoecology 159:113–143.

Morrow, J. E., and Gnecco, C. (eds.) (2009). Paleoindian Archaeology: A Hemispheric Perspective,

University Press of Florida, Gainesville.

Newby, P., Bradley, J., Spiess, A., Shuman, B., and Leduc, P. (2005). A Paleoindian response to Younger

Dryas climate change. Quaternary Science Reviews 24: 141–154.

Pinter, N., and Ishman, S. E. (2008). Impacts, mega-tsunami, and other extraordinary claims. GSA Today18: 37–38.

Pitblado, B. L. (2003). Late Paleoindian Occupation of the Southern Rocky Mountains, University Press

of Colorado, Niwot.

Powell, J. F. (2005). The First Americans: Race, Evolution, and the Origin of Native Americans,

Cambridge University Press, Cambridge.

Renfrew, C. (ed.) (2000). America Past, America Present: Genes and Languages in the Americas andBeyond, McDonald Institute for Archaeological Research, Cambridge.

Roberts, R. G., Flannery, T. F., Ayliffe, L. K., Yoshida, H., Olley, J. M., Prideaux, G. J., Laslett, G. M.,

Baynes, A., Smith, M. A., Jones, R., and Smith, B. L. (2001). New ages for the last Australian

megafauna: Continent-wide extinction about 46,000 years ago. Science 292: 1888–1892.

Robinson, G. S., Burney, L. P., and Bruney, D. A. (2005). Landscape paleoecology and megafaunal

extinction in southeastern New York state. Ecological Monographs 75: 259–315.

Slaughter, R., and Skulan, J. (2001). Did human hunting cause mass extinction? (Response to Alroy, J.,

2001). Science 294: 1460–1461.

Steadman, D. W., Martin, P. S., MacPhee, R. D. E., Jull, A. J., McDonald, H. G., Woods, C. A., Iturralde-

Vinent, M., Hodgins, G. W. L., and Dickinson, W. R. (2005). Asynchronous extinction of Late

Quaternary sloths on continents and islands. Proceedings of the National Academy of Sciences 102:11763–11768.

Stuart, A. J. (1999). Late Pleistocene megafaunal extinctions: A European perspective. In MacPhee, R. D.

(ed.), Extinctions in Near Time: Causes, Contexts and Consequences, Kluwer Academic/Plenum

Publishers, New York, pp. 257–270.

Stuart, A. J. (2005). The extinction of woolly mammoth (Mammuthus primigenius) and straight-tusked

elephant (Paleoloxodon antiques) in Europe. Quaternary International 126–128: 171–177.

Stuart, A. J., Sulerzhitsky, L. D., Orlova, L. A., Kuzmin, Y. V., and Lister, A. M. (2002). The latest

woolly mammoths (Mammuthus primigenius Blumenbach) in Europe and Asia: A review of the

current evidence. Quaternary Science Reviews 21: 1559–1569.

Surovell, T. A., and Waguespack, N. M. (2008). How many elephant kills are 14? Clovis mammoth and

mastodon kills in context. Quaternary International 191: 82–97.

Surovell, T. A., Waguespack, N., and Brantingham, P. J. (2005). Global archaeological evidence for

proboscidean overkill. Proceedings of the National Academy of Sciences 102: 6231–6236.

Surovell, T. A., Holliday, V. T., Gingerich, J. A. M., Ketron, C., Haynes, Jr., C. V., Hilman, I., Wagner,

D. P., Johnson, E., and Claeys, P. (2009). An independent evaluation of the Younger Dryas

extraterrestrial impact hypothesis. Proceedings of the National Academy of Sciences 106:18155–18158.

Sykes, B. (1999). The Human Inheritance: Genes, Languages, and Evolution, Oxford University Press,

Oxford.

Tankersley, K. B. (2002). In Search of Ice Age Americans, Gibbs Smith, Salt Lake City.

374 J Archaeol Res (2011) 19:327–375

123

Page 49: fulltext (2)

Thomas, D. H. (2001). Skull Wars: Kennewick Man, Archaeology, and the Battle for Native AmericanIdentity, Basic Books, New York.

Ubelaker, D. H., Sturtevant, W. C., and Stanford, D. (eds.) (2007). Handbook of North American Indians,Volume 3: Environment, Origins, and Population, Smithsonian Institution Press, Washington, DC.

Walker, R. B., and Driskell, B. N. (eds.) (2007). Foragers of the Terminal Pleistocene in North America,

University of Nebraska Press, Lincoln.

Webb, S. D. (ed.) (2006). First Floridians and Last Mastodons: The Page-Ladson Site in the AucillaRiver, Springer, Dordrecht, The Netherlands.

Wroe, S., Field, J., and Grayson, D. K. (2006). Megafaunal extinction: Climate, humans, and

assumptions. Trends in Ecology and Evolution 21: 31–62.

Yang, Z. Q., Verbeeck, J., Schryers, D., Tarcea, N., Popp, J., and Rosler, W. (2008). TEM and Raman

characterization of diamond micro- and nanostructures in carbon spherules from upper soils.

Diamond and Related Materials 17: 937–943.

Yang, Z., Schryvers, D., Rosler, W., Tarcea, N., and Popp, J. (2008). Micro- and nano-diamond particles

in carbon spherules found in soil samples. In Richter, S., and Schwedt, A. (eds.), EMC 2008, Vol. 2:Materials Science, Springer, Berlin, pp. 833–834.

Yesner, D. R., Barton, M. C., Clark, G. A., and Pearson, G. A. (2004). Peopling of the Americas and

continental colonization: A millennial perspective. In Barton, C. M., Clark, G. A., Yesner, D. R., and

Pearson, G. A. (eds.), The Settlement of the American Continents: A Multidisciplinary Approach toHuman Biogeography, University of Arizona Press, Tucson, pp. 196–213.

J Archaeol Res (2011) 19:327–375 375

123