Você está na página 1de 65

P1: FMN

Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

Journal of Archaeological Research, Vol. 8, No. 1, 2000

The Peopling of the New World: Present Evidence,


New Theories, and Future Directions
Stuart J. Fiedel1

The prevailing archaeological consensus on Paleoindian origins and colonization


of the Americas has been shaken by recent wide acknowledgment of pre-Clovis
occupation at Monte Verde, Chile, and by claims that ostensibly non-Mongoloid
skeletal remains might represent a precursor population. Recent mitochondrial
DNA studies have been interpreted by some as indicating an earlier and more
complex peopling of the continent. This paper reviews the current archaeological
and biological evidence, in America and northern Asia, for the origins of Native
Americans, assesses models of the colonization process in the light of new data
and a revised chronology, and suggests avenues for future research.
KEY WORDS: Paleoindian; Clovis; migration routes; peopling; colonization; Late Pleistocene;
America; Northeast Asia.

INTRODUCTION: A BRIEF HISTORICAL OVERVIEW


OF PALEOINDIAN THEORIES AND RESEARCH

In 1590, the Spanish cleric Jose de Acosta theorized that Native Americans
were descended from ancient savage hunters who had followed game animals
over a hypothetical land bridge from northeastern Asia into northwestern America
(Acosta, 1604, pp. 45, 57). Edward Brerewood, an English scholar, observed in
1614 that the people of America, lacking the civilized arts of China, India, and
Japan, resemble the old and rude Tartars [i.e., Mongols], above all the nations of
the earth. Northeast Asia, the Tartars homeland, if not continent with the west
side of America could only be separated by some narrow channell of the Ocean
(Brerewood, 1622, pp. 9697). Ever since these prescient early assessments of
the origins of Native Americans, the migration of their northeast Asian ancestors
across a land bridge, subsequently inundated by the Bering Sea, has been assumed
1 John Milner Associates, 5250 Cherokee Avenue, Suite 410, Alexandria, Virginia 22312.

39
1059-0161/00/0300-0039$18.00/0
C 2000 Plenum Publishing Corporation
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

40 Fiedel

by most serious scholars. The main dispute has been the date of their crossing (or
crossings) from Asia. In 1590, when no educated European thought the world itself
was more than 6000 years old, Acosta surmised that the migration had occurred
some 2000 years before his time.
Spurred by the acceptance, after 1859, of a greatly extended time frame for
human biological and cultural evolution in the Old World, and by the discovery
of early human remains in Europe, scholars in North and South America sought
comparable traces of an ancient presence in this hemisphere (Meltzer, 1991, 1994).
However, by the end of the century, it was clear that the superficial similarity of
some crude-looking American stone artifacts to Old World Paleolithic specimens
was no proof of the formers great antiquity. A conservative reaction set in, so that
all claims for ages of more than about 5000 years for artifacts or human skeletons
were dismissed outright or regarded skeptically.
But when obviously man-made spearpoints were found embedded within the
skeletons of extinct giant bison, near Folsom, New Mexico, in 1926, American
archaeologists and physical anthropologists were compelled to rethink their basic
assumptions about the prehistory of American Indians. The Folsom find, examined
in situ by prominent visiting scholars in 1927 and 1928, demonstrated the coexis-
tence of humans (Paleoindians) and giant mammals (megafauna) that died out
at the end of the Ice Age, then estimated as about 10,000 years ago.
Within the next decade, similar, but not identical points were found alongside
the bones of mammoths at Dent, Colorado, and at Blackwater Draw, near Clovis,
New Mexico (Cotter, 1937). Later investigations at Blackwater Draw in 19491950
demonstrated that these points and the associated mammoth skeletons occurred
in sediments stratified below the level containing Folsom points (Sellards, 1952).
Clovis-like fluted points were soon found across the whole of North America, and
a few even turned up in Central America. At the farthest southern tip of South
America, Junius Bird (1938) excavated points with fishtail-like stems, in apparent
association with bones of extinct horses. Very similar stemmed points were later
found in Ecuador and in Central America. These points were credible stylistic
descendants of the Clovis type.
Application of the new radiocarbon dating method in the early 1950s initially
yielded problematic late dates for Folsom, but a date of 10,780 135 rcbp for the
Lindenmeier site, published in 1960, appeared valid. In 1959, the Lehner Clovis
site was dated at 11,290 500 and 11,180 140 rcbp (Haury et al., 1959; Haynes,
1992). Dates for other sites supported this chronology. [NB: rcbp refers to un-
calibrated conventional radiocarbon age; B.P. connotes corrected dates (Fiedel,
1999).]
So, by the mid-1960s, a coherent picture of initial human colonization seemed
to be emerging, as outlined by C. Vance Haynes (1966). Clovis artifacts were made
by the first inhabitants of the continent. Their ancestors, hunting people of the
northern Eurasian grasslands, had crossed the 1600-km-wide land bridge, called
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 41

Beringia, exposed by lowered sea level during the Late Pleistocene. However, at
its eastern end in Alaska, Beringia was sealed off from North America by the
coalescence of two massive ice sheets (the Cordilleran to the west, the Laurentide
to the east). As climate warmed toward the end of the Pleistocene, the glaciers
receded, and a passage opened between themthe ice-free corridor. Ancestral
Paleoindians ventured south through the corridor around 11,500 rcbp, stumbling
upon a hunters paradise of megaherbivores that had no prior experience of human
predation. Paul S. Martin (1973) seized upon this aspect of the Clovis-first theory,
creating an elegant model that explained both the ubiquitous appearance of fluted
spearpoints and the apparently simultaneous demise of numerous species of giant
Pleistocene fauna. Martin theorized that Paleoindians had engaged in a killing
spree, a blitzkrieg-like rapid advance that resulted in overkill and extinction of the
megafauna by 10,000 rcbp.
The Clovis and Folsom cultures had a very sophisticated lithic technology,
equivalent to the best stonework of the European Upper Paleolithic. However,
Old World toolkits are dominated by retouched blades and only rarely include
bifaces; indeed, bifaces have sometimes been regarded as holdovers from earlier,
more primitive Mousterian traditions. The American Paleoindian industry, focused
on production of bifacially chipped points, has been characterized as a sort of
advanced Mousterian (Muller-Beck, 1966), but this view was based in part on a
misunderstanding of its technology. The use of blades, struck from prismatic cores,
in Clovis toolkits has been known since a blade cache was discovered at Blackwater
Draw in 1962 (Green, 1963). More recent finds in the Southeast and Midwest (e.g.,
Sanders, 1990; Stanford, 1991) and reanalysis of Southwestern Clovis assemblages
(Goebel et al., 1991) have demonstrated the importance of retouched blade tools.
The theoretical implications of blade technology can best be understood in the
context of Old World cultural evolution. Simple stone toolsOldowan choppers
and flakeswere being made in East Africa more than 2.6 million years ago. This
basic lithic tradition was carried into eastern Asia by the earliest hominids who
migrated thereperhaps 2 million, perhaps 1 million years ago. About 1.5 million
years ago, African hominids, presumably of the Homo erectus or Homo ergaster
lineage, were making handaxes. It was once thought that this technology had spread
only to Europe and the Middle East, as far as India, but no farther into eastern Asia.
However, in the past few decades, handaxes have been reported from Mongolia,
Siberia, Korea, and China (Yi and Clark, 1983), and crude specimens are found
in Southeast Asia as well (in the Pacitanian industry). Nevertheless, Oldowan-like
choppers seem to have persisted for many millennia as the dominant lithic industry
in southeastern Asia, perhaps because more specialized tools were made of per-
ishable materials such as bamboo (Pope, 1984). Chipping of flakes from specially
prepared cores (using the Levallois or Mousterian techniques) was an innovation
of western hominids (probably archaic forms of H. sapiens) about 200,000 B.P.
Levallois-Mousterian technology had spread as far eastward as the Altai region of
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

42 Fiedel

Siberia before 43,000 rcbp. This Middle Paleolithic method of flaking from pre-
pared cores was a precursor of the techniques later used to strike long, thin flakes
(blades) from conical or prismatic cores. The first blade-based assemblages proba-
bly appeared more than 70,000 years ago in Africa. The blade-based Aurignacian
industry replaced the Mousterian in the Near East and Europe between 40,000
and 32,000 rcbp, and similar assemblages were present even earlier in Siberia
(Goebel and Aksenov, 1995). It is important to note that the first significant use
of bone and antler for tools is associated with blade industries; this may represent
some sort of cognitive breakthrough (Mithen, 1996). In Europe, the appearance
of blade-based assemblages coincides with the replacement of local Neanderthals
by anatomically modern Homo sapiens. It is likely that the Upper Paleolithic of
Siberia also marks the arrival of modern humans in the region, presumably replac-
ing the anatomically archaic inhabitants, although some Russian archaeologists
(e.g., Derevianko et al., 1998) see evidence of a gradual technological evolution
from the local Middle Paleolithic, which would be inconsistent with such complete
biological replacement.
Jesse Jennings (1974, p. 72) characterized the then-known Siberian Paleo-
lithic material as clearly derivative from the Afro-European Paleolithic cultures,
while the Asiatic Pacific Coast remains are seen as belonging to the ancient
Southeast Asian tradition long regarded as different from the European assem-
blage. He believed that the earliest American artifacts originated in this coastal
Asian tradition, yet noted the lack of Asiatic prototypes for the beautifully fluted
and pressure-flaked American point and blade types (p. 74). Similarly, Gordon
Willey (1966) suggested that, if people came to America more than 30,000 or
40,000 years ago, they brought a simple tool kit with no blades or points, in
the tradition of the ancient Lower or Middle Paleolithic Chopper-Chopping Tool
Industry of southeast Asia. Although he admitted that there were no thoroughly
acceptable associations of manmade artifacts, extremely early radiocarbon dates,
and convincing middle or early Wisconsin geological contexts, Willey still felt
that there remains strong suggestive evidence of a very early migration, in the
form of crude chopper-chopping tools found in North and South America. Alex
Krieger (1964) argued that these artifacts, although generally found in surface
contexts of indefinite date, could be assigned to an early pre-projectile stage.
Yet, as Jennings observed, the surface finds of American chopper-scrapers meet
none of the minimum control standards archeologists insist upon for any ancient
complex if it is to be deemed truly ancient (1974, p. 75), while the few stratified
occurrences (e.g., at the C. W. Harris site in California and the South Yale site in
British Columbia) were only of early Holocene age.
In 1974, Willey cited new finds in the Yukon (Old Crow) and in Mexico
(Valsequillo and Tlapacoya) as providing good evidence that man has been in the
New World for as much as 25,000 to 30,000 years if not longer. Jennings (1974,
p. 76) expressed disappointment that no convincing proof of the antiquity of the
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 43

pre-projectile stage had yet emerged, but in 1983, buoyed by evidence from Old
Crow, Shriver, and Meadowcroft of pre-11,000 occupation, he was more optimistic:
Further research should continue to push the evidence of human occupancy of
both American continents farther back into time, perhaps into the mid-Wisconsin
(40,000 B.P.) range (Jennings, 1983, p. 63). The eager acceptance by these doyens
of American archaeology of apparent evidence of pre-Clovis occupation shows that
the scientific establishment has not been predisposed to crush such claims. The
problem for pre-Clovis advocates has been one of cumulative negative evidence,
not paradigmatic rigidity (Meighan, 1983). In contrast to the optimism of Jennings
and Willey about pre-Clovis finds, Waters (1985) stated that after an evaluation
of the evidence, collected by numerous researchers over many decades, the Clovis
Culture still remains the oldest unequivocal evidence for man in the Americas south
of the former continental ice sheets. I reached the same conclusion in 1987: It
is only after about 9500 B.C. that we find entirely convincing evidence of human
occupation in North America (Fiedel, 1987, p. 56), although I had to admit that
the archaeological record offers tantalizing hints of earlier occupation (p. 51).
If the presence of ca. 30,000-year-old or older simple lithic industries were
proven, major problems of cultural evolution would confront American archaeol-
ogists. If crude pebble choppers were sufficient equipment for American foragers
for some twenty millennia or more, what circumstances would have provoked the
sudden transformation from Lower to Upper Paleolithic toolmaking (with no in-
tervening Middle Paleolithic stage), across the entire hemisphere, in the course
of only a few hundred years at 13,000 B.P.? If the creation of Clovis points and
associated stone tools such as blades, endscrapers, sidescrapers, and gravers, and
bone and ivory foreshafts was the result of innovation after long isolation from the
Old World, are Paleoindian similarities to the European Upper Paleolithic (e.g.,
endscrapers, blades, red ocher, bone wrenches, beveled and scored bone points)
(Haynes 1980a, 1982; Stanford, 1991) the product of accidental convergence?
This seems unlikely (Cotter, 1981), but archaeologists accept even more specific
convergences at a Neolithic technological level (e.g., grooved bark beaters, cord
impression of clay pots, copper axes) as coincidental similarities, so a historical
relationship is not necessarily implied. In fact, a recent renewal of interest in the
superficial resemblance of Solutrean and Paleoindian bifaces (e.g., Stanford, 1998)
has shown where a facile interpretation of such convergence can lead [for a more
cautious assessment of the similarities, see Bordes (1968, p. 217)]. Solutrean flint
knappers created beautiful leaf-shaped bifaces, using techniques that seem to have
been replicated by Clovis and later Plano toolmakers. But the Solutrean indus-
try, prevalent only in France and the Iberian Peninsula, did not persist beyond
ca. 16,500 rcbp (19,000 B.P.)about 6000 years before the appearance of Clovis.
Besides, the same biface-thinning techniques were practiced by Russian Upper
Paleolithic toolmakers (Bradley, 1997) who can be more credibly connected with
the Asian ancestors of the Paleoindians.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

44 Fiedel

Even some researchers who accept the reality of a very early human pres-
ence in the Americas cannot quite bring themselves to postulate an in situ rapid
evolution from pebble tools to Clovis. Butzer (1991, p. 145) must suppose two
Pleistocene penetrations, one before 30,000, the other after 15,000 B.P. When
the expansive Paleoindian populations and their descendants subsequently dis-
persed through the Americas, they settled some regions for the first time and, in
others, they would have swamped any low-density, older occupants. It would be
naive to expect tangible linguistic or biological traces of such mid-Wisconsinan
immigrants in the modern indigenous population . . . any more than Neanderthal
lexical or dental components can be expected in the human mosaic of contem-
porary Europe (Butzer, 1991, p. 152). Alan Bryan (1978, 1983, 1991), who has
long advocated an initial migration more than 30,000 years ago, is more sanguine
about convergent multilinear evolution. Indeed, he went so far as to suggest that
biological evolution probably occurred within America from a paleoanthropine
ancestor to explain at least some of the highly diverse populations of modern
American Indians (Bryan, 1978, p. 323). Bryan has theorized that, in several
areas, generalized foragers carrying a simple core/flake lithic tradition indepen-
dently innovated specialized hunting and a number of regionally distinct bifacial
spearpoints (Clovis east of the Rockies, stemmed points in the West, fishtails in
southern South America, tanged points in Brazil, and leaf-shaped El Jobo points
in northern South America). A major difficulty for this view is that there are no
transitional industries that resemble, or are even taxonomically equivalent to, the
Old World Middle Paleolithic. If ancient American cultures progressed on an evo-
lutionary trajectory parallel to that of the Old World, they somehow skipped a
whole lithic stage, jumping from Lower to Upper Paleolithic technology in a few
centuries.

PRE-CLOVIS OCCUPATION: THE CURRENT STATE


OF THE EVIDENCE

Over the years, the claims of great antiquity for dozens of sites have failed
to withstand critical scrutiny (e.g., Haynes, 1974; Lynch, 1990; Waters, 1985).
The list of plausible pre-Clovis contenders has now shrunk to a handful of sites:
Meadowcroft Rockshelter, Cactus Hill, Topper, Wilson Butte Cave, Fort Rock
Cave, the Chesrow complex, Monte Verde, Bluefish Caves, Taima-Taima, Pedra
Furada, and Valsequillo (Figs. 1 and 2).
Meadowcroft Rockshelter. A lithic assemblage consisting of small blades, a
flake knife, and a reworked lanceolate point ascribed to the Miller type was re-
covered from Stratum IIA of this rockshelter, located in southwestern Pennsylvania
(Adovasio et al., 1990). Radiocarbon dates associated with this material ranged
from 19,000 to 11,000 rcbp (in appropriate stratigraphic order). There is no
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 45

Fig. 1. Paleoindian sites in North America: (1) Batza Tena (obsidian fluted points); (2) Mesa;
(3) Nenana sites; (4) Broken Mammoth and Swan Point; (5) Old Crow and Bluefish Cave;
(6) Prince of Wales Island; (7) Queen Charlotte Islands; (8) Namu; (9) Charlie Lake Cave;
(10) Vermilion Lakes; (11) Manis; (12) Marmes; (13) Fort Rock Cave; (14) Buhl; (15) Wilson
Butte; (16) Anzick; (17) Agate Basin; (18) Jim Pitts; (19) Lindenmeier; (20) Dent; (21) Folsom;
(22) Blackwater Draw; (23) Murray Springs, Naco Lehner; (24) Aubrey; (25) Daisy Cave,
Santa Rosa Island; (26) Santa Isabel Iztapan; (27) Tlapacoya; (28) Valsequillo; (29) Big Eddy;
(30) Bostrom; (31) Adams; (32) Carson-Conn-Short; (33) Dust Cave; (34) Chesrow; (35) Gainey;
(36) Paleo Crossing; (37) Lamb; (38) Meadowcroft; (39) Shawnee-Minisink; (40) Bull Brook;
(41) Vail; (42) Debert; (43) Topper, Big Pine Tree; (44) Page-Ladson; (45) Little Salt Spring. Stip-
pled areas depict ice sheets ca. 12,000 rcbp; dashed line is approximate shoreline at 12,000 rcbp.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

46 Fiedel

Fig. 2. Paleoindian sites in Central and South America: (1) Turrialba (fishtail and Clovis-like
points); (2) Madden Lake (fishtail points); (3) El Inga; (4) Quebrada Jaguay; (5) Quebrada Tacahuay;
(6) Monte Verde; (7) Cueva del Medio and Cueva del Lago Sofia; (8) Fells Cave; (9) Los Toldos;
(10) Piedra Museo; (11) Cerro La China; (12) Lagoa Santa and Lapa Vermelha; (13) Santana do
Riacho; (14) Lapa do Boquete; (15) Toca da Esperanca; (16) Pedra Furada and Sitio do Meio;
(17) Pedra Pintada (Monte Alegre); (18) El Jobo; (19) Taima Taima and Muaco.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 47

question that the site was painstakingly excavated, although detailed plan views
and profiles of features and artifact distributions have yet to be published. The
problems here are that (1) the plant remains, the snail shells, and the few animal
bones found in Stratum IIA all belong to species found in deciduous forests, not
the boreal spruce/pine forest with patches of tundra that prevailed in this region
just after the glacial maximum, when the ice sheet lay 80 km to the north (Mead,
1980; Wright, 1991); (2) there is some doubt whether the dated material from IIA
was human-created charcoal (Haynes, 1980b, 1991a); (3) there are no Clovis or
early Early Archaic components with diagnostic artifacts; and (4) the Miller point
type, as reconstructed from the Meadowcroft specimen and others found at nearby
open sites, looks like an Agate Basin variant (see Adovasio, 1993, p. 212). This
suggests an age of about 10,000 rcbp, not 12,80011,300 rcbp as indicated by
radiocarbon assays. The ca. 15,000 rcbp dates from Stratum IIA are probably at
least 2000 to 3000 years too old.
Cactus Hill. Within a ridge of stabilized wind-blown sand in southeastern
Virginia, Archaic assemblages occur in expected stratigraphic order. Below them
lies a Clovis horizon, with a date of 10,920 250 rcbp. Lying several inches below
the Clovis level is a small assemblage of quartzite blades and bladelets, and two
pentagonal, unfluted points made of a metavolcanic stone (Johnson, 1997; McAvoy,
1997). Dates of 15,070 70 and 16,670 730 rcbp (but also an anomalously late
date of ca. 9250 rcbp) have been obtained on charcoal particles from this level. The
problematic issues at this site are twofold. (1) If the ca. 15,000 rcbp (18,000 B.P.)
date is correct, some 5000 years intervened between the pre-Clovis and the Clovis
occupations, while only 3 in. of sand accumulated. (2) Downward drifting of
cultural material and animal burrowing are acknowledged problems at the site,
and these processes are blamed for the 9250 rcbp date and several others that seem
too recent. The possibility that the dated charcoal represents prehuman, naturally
produced material, e.g., from forest fires, cannot yet be excluded.
Topper. Like Cactus Hill, this site near the Allendale chert source in South
Carolina contains a rather sparse lithic component stratified below (in this case,
about a meter below) a later Paleoindian (Dalton) lithic assemblage (Goodyear
et al., 1998). The deepest material could be Clovis age, although the depth of
intervening sediments is suggestive of a longer hiatus than the ca. 1000 years
that separate Clovis from Dalton. Recent radiocarbon dates on dispersed charcoal
from the lower level have been disappointingly recent (D. Anderson, personal
communication, 1999).
The Chesrow Complex. In southeastern Wisconsin, near Kenosha, artifacts
made of local lithic material had been thought to represent a late Paleoindian
variant. However, a few flakes were found in association with a mammoth at
the Schaefer site, now dated by three radiocarbon assays (one on purified bone
collagen, two on spruce wood) to about 12,300 rcbp. At the nearby Hebior site,
a rather crude biface was found lying within a pile of mammoth bones. Other
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

48 Fiedel

mammoths have been found in this area, dating to ca. 13,450 rcbp, but butchering
by humans has yet to be conclusively demonstrated (Overstreet and Stafford, 1997).
Bluefish Caves. In the northern Yukon, a few stone artifactswedge-shaped
cores, microblades, and burinswere found in apparent association with remains
of mammoth, horse, bison, and other Pleistocene mammals (Morlan and Cinq-
Mars, 1982). Radiocarbon dates on the bones ranged between 12,210 210 and
25,000 rcbp. If the dates and association are valid, this would be the oldest cultural
material from eastern Beringia. However, it remains uncertain whether the artifacts,
typical of the somewhat later Paleoarctic tradition (11,700 rcbp and later), are
displaced.
Valsequillo. At this location in Puebla, Mexico, four sites were found within
a 30-m-thick gravel formation on a peninsula in a reservoir (Irwin-Williams, 1967,
1978, 1981; Waters, 1985). In descending stratigraphic order, they are called Huey-
atlaco, Tecacaxco, El Mirador, and El Horno. A fifth locality, Caulapan, yielded a
radiocarbon date of 21,850 850 rcbp for freshwater gastropods associated with a
crudely worked artifact (possibly nonartifactual). Such organisms are notoriously
unreliable as carbon samples. The upper component at Hueyatlaco (stratified 10 m
below the top of the gravels) contained a bifacial point, a stemmed point, and
knives, scrapers, burins, and utilized flakes. Similar assemblages have been dated
to about 9000 to 11,000 rcbp elsewhere in Mexico [e.g., in Tamaulipas and Puebla
(MacNeish, 1983; Wormington, 1957, p. 202)]. A lower component consisted of
seven unifacially retouched blades, found in close association with bones of extinct
mammals. A few similar blades were found at Tecacaxco and El Mirador, also with
extinct fauna. At El Horno, at the base of the gravel formation, 13 flakes, with no
blades, occurred in association with late Pleistocene fauna. The dates obtained by
experimental fission-track and uraniumthorium techniques unfortunately did little
to clarify the situation (Steen-McIntyre et al., 1981; Szabo et al., 1969); a proposed
age of ca. 250,000 years was unacceptable to the excavator (Irwin-Williams, 1981)
and even to the most ardent advocates of an early human presence (e.g., Bryan,
1978, p. 315). Despite Waters (1985, pp. 133, 137) assessment that Valsequillo
was the most convincing of the putative sites in the 27,000 to 11,500 rcbp range,
it has somehow dropped out of the literature.
Taima-Taima (Bryan et al., 1978). At this Venezuelan site, an El Jobo point
was found lying (but not actually embedded) within the pubic cavity of a juvenile
mastodon. Two other point fragments, a scraper, and a flake also were found. The
artifacts and mastodon bones, possibly indicative of butchering, lay within a gray
sand layer, interpreted as a spring deposit. Radiocarbon dates were obtained for
samples of soil organics, bone, and twigs interpreted as the stomach contents of
a mastodon. Dates for the soil ranged from 12,580 150 to 13,390 130 rcbp,
with some stratigraphic reversals. Collagen samples from mastodon bones came
out at 13,010 280 and 14,440 435. Dates for the supposed stomach contents
ranged from 12,980 85 to 14,200 300. Another wood sample yielded a date of
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 49

11,860 130. The presence of coal in the vicinity, and permeation of the deposits
by groundwater, raised the possibility that the early dates reflected contamination
with old carbon (Haynes, 1974). In any case, it is fairly certain that this material
is earlier than ca. 10,300 rcbp (the age obtained for an overlying layer of clay)
(Waters, 1985). The dates of ca. 12,00014,000 rcbp have become more credible
because of the acceptance of ca. 12,00012,700 rcbp dates associated with vaguely
similar bifaces at Monte Verde. El Jobo points also were found in association with
mastodon bones at the Muaco and Cucuruchu sites in Venezuela, but evidence of
disturbance by spring waters precludes uncritical acceptance of the associations
or the dates obtained (Dincauze, 1984; Lynch, 1990).
Monte Verde. Monte Verde, in the lake district of Chile, was dramatically
promoted from troublesome enigma to paradigm buster by the announcement in
1997 of the consensus reached by a small archaeological delegation that viewed the
remnants of the now largely obliterated site. Most significant was the conversion
of Dena Dincauze and Vance Haynes, both previously skeptical of all pre-Clovis
claims. The trip to Monte Verde was arranged as a latter-day counterpart of the 1927
Folsom site inspection that resulted in consensual acceptance of a Late Pleistocene
human presence in North America. It should be noted, however, that at Folsom the
crucial points still lay in situ amid the bison bones when the visiting specialists
arrived; at Monte Verde, the excavations ceased in 1985, and no artifacts were
viewed in their original positions by the visitors in 1997. Furthermore, the social
context of American archaeology has changed a good deal in the intervening seven
decades (see Meltzer, 1994, p. 18). Professional archaeologists now number several
thousand, not several dozen, and they tend be averse to unquestioning acceptance
of the opinions of authoritative experts. So the new consensus has not completely
silenced skeptics.
Tom Dillehay, who excavated the site and supervised exhaustive specialist
analyses of the artifacts and floral and faunal remains, has made a convincing case
that Monte Verde was a human settlement, and that the radiocarbon dates have
not been affected by any contaminants. Most compelling are a childs footprint,
strands of fiber knotted around wooden tent pegs, and clearly planed logs. These
logs formed a complex of 12 connected rectangular rooms, probably roofed with
mastodon hides (pieces of which have been identified, adhering to some logs).
A separate ovoid structure may have served as a shamans residence. Preserved
organic remains include many species of plants, among them wild potatoes, medic-
inal plants (chewed as quids), and salt-rich seaweed carried to the site from the
Pacific coast, some 30 km to the west. The plants were processed using wooden
mortars and grinding stones. Hunting (or perhaps opportunistic scavenging) is at-
tested by the presence of six mastodon-like proboscideans (gomphotheres), as well
as an extinct species of guanaco. The Monte Verde lithic assemblage is strangely
depauperate, however. A pressure-flaked, narrow lanceolate point and two similar
midsection fragments represent a very proficient stone-chipping tradition. A slate
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

50 Fiedel

rod also was skillfully produced by grinding. On the other hand, several hundred
supposed tools are merely unmodified but possibly utilized broken river cobbles.
Recognizable debitage from lithic flakinga practically universal marker of Stone
Age human encampmentsseems, inexplicably, to be almost absent (only 14 ev-
ident flakes are catalogued). Three stone scrapers, mounted in wooden hafts with
bitumen, were described and illustrated in a 1988 article about the site (Dillehay,
1988), but only one is described in the massive final report on the archaeological
investigations (Dillehay, 1997).
Monte Verde appears to represent a single human occupation of perhaps
several seasons duration (Dillehay, 1997). The associated radiocarbon dates for
wood, bone, and ivory artifacts range from ca. 11,700 to 12,800 rcbp. There is an
obviously erratic outlier of 13,565 250 rcbp (ca. 16,000 B.P.); Dillehay (1989)
formerly regarded this as the most reliable date for the settlement, but now (1997)
explains it as dating already old wood from a very long-lived larch (alerce) tree.
He now dates the occupation to around 12,570 rcbp, based on a suite of six dates
falling between 12,230 140 and 12,780 240 rcbp. However, four radiocarbon
assays seem to imply a later date. A bone-derived date of 11,990 200 and a date
of 12,000 250 for an ivory artifact may justifiably be disregarded as too recent
because of the sample materials, but dates of 11,920 120 and 11,790 200 rcbp
for carbonized wood samples cannot be so easily dismissed. Dillehay (1997) now
surmises that these samples really belong to the peat layer and are secondarily
deposited in the occupation zone. The overlying peat that sealed and preserved
the organic remains immediately after abandonment of the settlement (Tuross and
Dillehay, 1995) dates between 10,300 and 12,000 rcbp, with four dates in the
11,600 to 11,800 rcbp range. The most likely age of the human occupation thus
appears to be only slightly earlier than the peat, or about 11,80012,000 rcbp
(ca. 14,10013,500 B.P.). Dating of some of the shorter-lived organic materials
preserved at the site (e.g., the chewed plant quids, the knotted reed twine, and
the mastodon skin fragments), which has not been attempted, might help to pin
down its age with greater certainty. The existing radiocarbon dates of ca. 12,700 to
11,800 rcbp at Monte Verde are consistent with the regional beetle evidence that
indicates the establishment of a closed southern beech forest after 13,000 rcbp,
and the logs and worked wood at the site also suggest a forest setting for the
occupation.
It must be added that there is possible evidence of a much earlier human
presence at Monte Verde. Some distance away from the ca. 12,000 rcbp deposits,
and in a lower (by 80 cm) and older layer, two dozen modified pebbles of basalt
and andesite were found near two ostensible hearths. Carbonized wood from these
features dates to 33,370 530 and more than 33,020 rcbp (about 35,000 B.P., dur-
ing a relatively warm interstadial period). Judging from published illustrations, at
least one of the stones has certainly been chipped by humans. This tool has yielded
protein residue that may be derived from a mastodon (Tuross and Dillehay, 1995).
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 51

It strains credulity to accept that two pioneering groups could, by coincidence,


have set up camps so close to one another, at so remote and improbable a location,
more than 20,000 years apart. Yet this is what we must conclude, if the dates are
correct and the association is valid. However, even Dillehay is not yet ready to ac-
cept these implications of Monte Verde I. New excavations are said to be planned
for 2001 (New York Times, 1998).
Monte Verde has profound implications (Meltzer, 1997). If the dates are
accurate, they push initial migration into the Americas back to at least 14,000 to
14,500 B.P. (12,00012,600 rcbp). As important as this is, it is more troubling that,
if the organic materials had not been preserved and if the handful of well-made
bifaces had not turned up, the lithic assemblage alone would not meet the long-
established criteria for recognition of human manufacture. We might be missing
hundreds of early sites because they consist of only expedient tools, indistinguish-
able from other stream-rounded and broken cobbles.
Even setting aside the ambiguous Monte Verde I component, Monte Verde II
remains unique at its ostensibly early date. The problems with samples, contexts,
and associations that preclude acceptance of pre-11,500 rcbp radiocarbon dates
from other South American sites (e.g., El Abra, Quirihuac, Muaco, Taima Taima)
have been amply discussed by Lynch (1990; see also Waters, 1985).
Pedra Furada. At this rockshelter site in northeastern Brazil, French archae-
ologist Ni`ede Guidon has excavated multiple layers of charcoal concentrations,
interpreted as hearths, and broken quartz and quartzite cobbles that are said to be
pebble tools (Guidon and Delibrias, 1986). The hearths yielded dates ranging
from 17,000 to more than 40,000 years, in stratigraphic order. There are paintings
on the shelter walls, and it is reported that fragments of the paintings were found in
ca. 17,000-year-old occupation levels. Upper levels of the site contain undoubted
tools made of exotic cherts. Several other sites in the same region have yielded
comparably early dates, but few details are available concerning the contexts. Like
Monte Verde, Pedra Furada was examined several years ago by a small party of
North American archaeologists (Meltzer et al., 1994). It did not fare as well. They
concluded that the broken cobbles were not artifacts and that the hearths were
not created by human activity.
Fort Rock Cave, Oregon. A date of 13,200 720 was obtained on charcoal
associated in the same arbitrary 10-cm level with two points, a mano fragment, and
11 flaked stone tools (Carlson, 1983, p. 76). One of the points is concave based; the
second a stubby, much resharpened stemmed point. A date of 10,200 230 rcbp
came from 20 cm higher. The temporal association of the charcoal with the points
is dubious, because the site was excavated in arbitrary levels (Waters, 1985).
Elsewhere, similar stemmed points have not been reliably dated as earlier than
10,700 rcbp (Faught, 1996; Thompson, 1985; Willig and Aikens, 1988), and ev-
idence from the Dietz site in Oregon suggests that they are post-Clovis. Both
stemmed and concave-based points occurred in the ca. 10,200 rcbp level.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

52 Fiedel

Wilson Butte Cave, Idaho. In Stratum C, a basalt knife, a blade, a utilized


flake, and two modified bones were found. A date of 14,500 500 rcbp was ob-
tained on apatite from rodent bones at the bottom of this stratum. A lower stratum,
E, produced a bone-based date of 15,000 800 rcbp. Dating of the apatite or in-
organic fraction of bone frequently yields dates that are much too old (Stafford,
1994; Taylor, 1992). For another example, consider the case of a clearly artifac-
tual serrated scraper, made on a caribou tibia, from the Old Crow basin in the
Yukon. It yielded a radiocarbon age of ca. 27,000 rcbp on the apatite fraction;
subsequent dating of the organic fraction put it at only 1350 150 rcbp. This re-
dating (Nelson et al., 1986) effectively quashed the theory, popular in the 1970s
and 1980s, that other broken bones from Old Crow represented a 30,000-year-old
nonlithic industry.

A REVISED CHRONOLOGY FOR CLOVIS

Based on initial radiocarbon results and Pleistocene stratigraphic research,


Clovis was usually ascribed an age of ca. 12,000 to 11,000 rcbp. In the early 1980s,
averaging of replicate radiocarbon dates on charcoal samples from Paleoindian
sites such as Lehner and Murray Springs resulted in a tighter estimate for the
florescence of Clovis culture: 11,200 to 10,900 rcbp (Haynes, 1992, 1993; Haynes
et al., 1984). Dates for Folsom sites sometimes overlapped with Clovis dates, in the
range of 10,600 to 11,000 rcbp, but since the Clovis mammoth kills were stratified
below Folsom material at Blackwater Draw, it seemed certain that these cultures
were successive, not contemporaneous [as Wilmsen (Wilmsen and Roberts, 1984)
had suggested].
Since the 1970s, archaeologists had realized that fluctuating ratios of 14 C and
12
C in the past had affected radiocarbon dates significantly, so that radiocarbon
ages were about 1000 years too young around 8000 B.P. Preliminary indications
suggested that this difference narrowed in the early Holocene, so Paleoindian spe-
cialists were not too concerned about the possible effects of calibration on their
chronology. However, recent research has shown that Late Pleistocene radiocarbon
dates are about 2000 years younger than calendrical ages (for a detailed discus-
sion, see Fiedel, 1999). Atmospheric fluctuations in carbon dioxide and radiocar-
bon caused a pattern of wiggles, sharp jumps, and level plateaus between 15,000
and 11,000 calendrical years B.P. Clovis-associated dates ranging from 11,300 to
10,700 rcbp correspond to a short period from about 13,100 to 12,900 B.P. in calen-
dar years (Taylor et al., 1996). The earliest Clovis dates of ca. 11,55011,600 rcbp,
obtained from the Aubrey site (Ferring, 1995), correspond to ca. 13,400 B.P. A
date of 11,900 80 rcbp was recently obtained for the Clovis component at the
Big Eddy site in Missouri, but other dates for the same stratum cluster between
11,400 and 10,900 rcbp (N. Lopinot, personal communication, 1998). Evidence of
a radiocarbon reversal in several records, such as Swiss lake bed cores, suggests
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 53

that some Clovis dates of about 10,70010,900 rcbp may be about 250 years earlier
than ostensibly coeval Folsom dates (i.e., ca. 13,150 vs. 12,900 B.P.).

NORTH AMERICA, 14,50013,500 B.P.: PRE-CLOVIS,


PROTO-CLOVIS, NON-CLOVIS?

If Monte Verde was occupied by hunter-foragers before 12,000 rcbp (ca.


14,100 to 13,500 B.P.), ancestors of this population must have traversed either
the coast or interior of North America at a still earlier date. However, no proven
sites of comparable or greater antiquity have been found in North America. Al-
though the differences between Clovis and Monte Verde lithic assemblages are
striking, Carlson (1991) has envisioned the possible development of Clovis tech-
nology from a Monte VerdeEl Jobo-like base. By assuming earlier migration of
bearers of the latter tradition through an ice-free corridor at about 12,500 rcbp
(14,700 B.P.), the basic dynamics of the Clovis-first model can be retained. This
would require the existence of a proto-Clovis population somewhere in North
America before 13,500 B.P., employing an Upper Paleolithic-style lithic toolkit
composed of bifaces and blades, but lacking fluted points (Carlson, 1991; Haynes,
1987).
A possible ancestor for this hypothetical culture can be found at Ushki I, in
Kamchatka, where Level 7, dated to about 14,000 rcbp (17,000 B.P.), contains
bifaces and blades (Fig. 3). Alone among the known Northeast Asian sites of this
age [e.g., the Dyuktai sites of the Aldan Basin (Mochanov and Fedoseeva, 1996a)],
Ushki 7 lacks microblades and wedge-shaped cores, which do appear in the over-
lying level (Dikov, 1996). Microblades also appear to be absent from Nenana
assemblages in central Alaska and from Clovis (Hoffecker et al., 1993). In both
Nenana and Clovis, many tools were made by retouching macroblades, a basic
similarity that has been cited to suggest that the Nenana complex was ancestral
to Clovis (Goebel et al., 1991; Yesner, 1998a). However, Nenana has only small
triangular points and teardrop-shaped (Chindadn) points, quite unlike the diagnos-
tic Clovis fluted points, and dates for Nenana sites have generally fallen around
11,200 rcbp (Pearson, 1997a), no earlier than Clovis. Recent research makes the
Nenana complex a little older. The earliest Nenana level at Broken Mammoth dates
to ca. 11,770 rcbp (Holmes, 1996), lending credence to a date of 11,820 200 rcbp
from Walker Road. These radiocarbon dates push the Nenana complex back to ca.
13,40013,900 B.P. The Dyuktai-derived Denali or Paleoarctic culture appears
contemporaneous with Nenana, on the basis of dates of ca. 11,660 rcbp associ-
ated with a microlithic assemblage at Swan Point (Holmes et al., 1996). These
dates have been particularly welcome news for Frederick West, who has long ar-
gued that the Paleoarctic industry, despite its microlithic character, was ancestral
to Clovis (West, 1981, 1996). Another early Alaskan culture is represented by
unfluted, Agate Basin-like points at the Mesa site. Most dates from this site are
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

54 Fiedel

Fig. 3. Paleolithic sites in Northeast Asia: (1) Afontova Gora; (2) Ust Kova; (3) Malta; (4) Varvarina
Gora; (5) Arta; (6) Avdeikha; (7) Diring Yuriakh; (8) Ikhine; (9) Ust Mil; (10) Verkhne Troitskaya;
(11) Dyuktai Cave; (12) Selemdga; (13) Khummi; (14) Gasya; (15) Ustinovka; (16) Berelekh;
(17) Bolshoi Elgakhchan; (18) Uptar; (19) Khukhtuy; (20) Ushki.

ca. 10,000 rcbp, but one hearth (perhaps a remnant of an earlier Paleoarctic occu-
pation) produced dates of 11,660 80 and 11,190 70 rcbp (Kunz and Reanier,
1994).
It thus appears that there were at least two distinct cultural traditions in Alaska
at the end of the Pleistocene. It may be that Clovis emerged from an Ushki or
Nenana-like antecedent in northern Alaska about 13,500 B.P. (11,600 rcbp). The
small fluted points found on the surface or in shallow mixed contexts in northern
Alaska unfortunately cannot be dated (Clark and Clark, 1983; Reanier, 1995).
They are usually considered to represent a northward reflux of relatively late
(ca. 10,600 rcbp?) Paleoindians from the northern Plains (Clark, 1991). A mi-
gration at that time would have been difficult, because the immigrants would have
confronted expanding Denali populations. Still, the Mesa finds appear to represent
a similar northern reflux of the Agate Basin complex ca. 10,000 rcbp, so an ear-
lier movement by late Clovis people would not be unique. On the other hand, the
geographic segregation of fluted point finds in northern Alaska from the Nenana
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 55

sites in central Alaska could indicate two contemporaneous but distinct cultures. If
identifications of mammoth residues on Alaskan points are valid (Loy and Dixon,
1998), the case for an Alaskan (or eastern Beringian) emergence of basic Clovis
lifeways would be strengthened. There is no evidence that Nenana people hunted
mammoth, which may already have been extinct in central Alaska. But if the
northern points are late, we must assume that Clovis ancestors did not practice
fluting when they first arrived south of the ice sheets. Perhaps, they invented fluted
points in order to hunt and butcher newly encountered mammoths (Haynes, 1987;
Pearson, 1997b).
A proto-Clovis assemblage might look like the material found at Santa Isabel
Iztapan, near Mexico City, in the early 1950s (Aveleyra, 1956). Both an unfluted
lanceolate point and a bipointed laurel leaf or Lerma-like point were found among
the bones of a single mammoth. This faunal association strongly suggests a pre-
Younger Dryas age, although the radiocarbon date reported in the 1950s was only
ca. 9250 rcbp (MacNeish, 1983). The lanceolate is a plausible prototype for Clovis
(Haynes, 1987), while the laurel leaf form, made by the same band of hunters, could
be a prototype for the South American complex represented by El Jobo and Monte
Verde bifaces.
Alternatively, proto-Clovis might be sought in the Southeast. It remains
to be demonstrated whether the two reported radiocarbon dates of ca. 15,000
16,000 rcbp (ca. 18,000 B.P.) accurately date the lithic material in the sub-Clovis
horizon at Cactus Hill, Virginia (McAvoy, 1997); this would be about 5000 calen-
dar years before Clovis. In any case, the small blades and conical cores at Cactus
Hill represent the sort of industry we might expect to find in a Beringia-derived
population, retaining some faint traces of an ancestral Dyuktai-like microblade tra-
dition. Bladelets also occur in a probably early Clovis assemblage at the Big Pine
Tree site in South Carolina (Goodyear, 1997). The long-controversial dates of ca.
15,000 and 13,000 rcbp for Stratum IIA at Meadowcroft Rockshelter (Adovasio
et al., 1990) remain problematic (Haynes, 1991a); however, the small blades found
there seem to fit the emerging picture of early Clovis or proto-Clovis toolkits in
the Southeast.
A few dates from other areas hint at a human presence around 12,000 rcbp
(ca. 13,50014,000 B.P.). In Florida, a wooden stake inserted into a tortoise shell,
presumably by human action, in the Little Salt Spring sinkhole, yielded a date
of 12,030 200 rcbp (but the shell itself is inexplicably much older, 13,450
190 rcbp) (Clausen et al., 1979, p. 611), and several dates for Zone D of the
Page-Ladson site, which contained an ostensibly incised mastodon tusk, ranged
from 13,130 200 to 11,770 90 rcbp (Faught, 1996, p. 162). In the absence
of a defined lithic industry, the affiliations of this dimly perceived occupation
remain obscure. Similarly, the early dates of 11,840 130, 11,700 95, and
11,450 110 rcbp for the pre-Folsom level at Agate Basin in Wyoming (Haynes
et al., 1984) cannot yet be tied to early Clovis or to any other culture. It should be
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

56 Fiedel

noted that dates for unfluted Goshen points in the northern Plains range from about
10,200 rcbp as far back as 11,570 170 rcbp (Frison, 1987, 1988, 1991, 1993,
1996), or ca. 13,400 B.P. If Goshen is not actually ancestral to Clovis, these early
dates could indicate that a still earlier common ancestral culture, without fluted
points, existed somewhere in the Plains before 13,400 B.P.
Initial Paleoindian emigration from Beringia probably occurred very soon af-
ter the beginning of the BllingAllerd period (14,700 B.P. ca. 12,500
12,900 rcbp), when temperatures rose abruptly in the northern hemisphere. In
western Beringia, summer temperatures may have been warmer than present
(Lozhkin et al., 1993). The Laurentide ice sheet suffered catastrophic ablation
ca. 12,300 rcbp, causing a global meltwater pulse in the ocean (Fairbanks et al.,
1992; Sowers and Bender, 1995). Although it may not yet have been inviting
for long-term habitation (Mandryk, 1996), the ice-free corridor would have been
passable at this time (Jackson and Duk-Rodkin, 1996). Ancestral Paleoindians
probably moved south in an opportunistic response to ameliorating climate, rapid
wasting of the Laurentide ice sheet, and widening and revegetation of the ice-free
corridor after 14,700 B.P.
Several groups of diverse Beringian ancestry may have undertaken explo-
ratory forays throughout North and South America after 14,700 B.P. However, in
the period of climatic oscillation after ca. 13,400 B.P. (11,400 rcbp), the people
who made Clovis points and associated tools perhaps enjoyed a unique techno-
logical/adaptive advantage (as efficient mammoth killers?) and rapidly expanded
their range throughout North America. In Central America, stylistic drift resulted
in the emergence of the fishtail point variant (probably derived from southeastern
Clovis) (Morrow and Morrow, 1997; Pearson, 1998). People who made fishtail
points reached Tierra del Fuego by about 13,000 B.P. If they encountered a pre-
cursor population in South America, of the Monte Verde or El Jobo culture, they
either absorbed or replaced them. Continuing stylistic drift and isolation, cou-
pled with rapid adaptation to diverse and changing early Holocene environments,
probably resulted in emergence of distinctive regional styles (e.g., Monte Alegre
in Brazil, Paijanense in coastal Peru, Dalton in the southern Midwest, etc.) by
ca. 12,500 B.P.

MIGRATION ROUTES: COAST OR INTERIOR?

The problem of identifying the initial human presence in the Americas has
always been linked to consideration of the geological and climatic preconditions for
migration. Migration from Asia would have been easiest when the Beringian land
bridge lay exposed above water, during periods of glacial advance and concomitant
lowered sea level. Contrary to previous interpretations, Beringia seems now to have
still been exposed until ca. 10,000 rcbp (Elias et al., 1996). In any case, even after
the land bridge was inundated, the distance from shore to shore across the Bering
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 57

Strait is only 90 km, and it is sometimes possible, in winter, to cross the frozen
strait by walking. Even when unfrozen, this distance would not have been very
challenging for people with simple watercraft; a comparable water barrier was
crossed by the first Australians more than 40,000 years ago, and people in Japan
were making sea voyages of more than 200 km before 17,000 rcbp (Yamaura,
1998). At its eastern edge, Beringia was sealed off from North America during
those periods when the mile-high Laurentide and Cordilleran ice sheets coalesced
in the Late Pleistocene. It has usually been assumed that the ancestral Paleoindians
were terrestrial hunters of big game, not littoral strandlopers. Thus, an interior
route from Alaska to the lower United States has been envisioned. The Paleoindians
would have had to thread their way through a 1200-km-long unglaciated passage,
the ice-free corridor between the ice sheets. Therefore, the timing of glacial
growth and retreat from the corridor is critical. As geologists now see it, the
corridor was sealed by coalescent ice sheets before 18,000 rcbp (21,000 B.P.).
The precise date of its opening is difficult to ascertain. Wright (1991) suggests
that the southern end had opened by about 12,000 rcbp. A vegetative cover of
herbs, grasses, sedges, mosses, and algae is reconstructed for the southern end of
the corridor at 12,000 to 11,700 rcbp. Meltzer (1995, p. 38; citing Burns, 1990)
gives a date of ca. 11,300 rcbp for the earliest mammals in the corridor. Mammoth,
horse, camel, elk, and musk-ox bones occur in Late Pleistocene terraces of the
Peace River Valley in Alberta, leading Fladmark et al. (1988, p. 382) to conclude
that the later stages of the ice-free corridor clearly were not inhospitable. This
characterization leaves open the question of habitability during the earlier stages,
before ca. 11,500 rcbp. I suspect that herbaceous vegetation rapidly colonized the
corridor soon after the sharp Blling warming at 12,600 rcbp (14,700 B.P.).
Fladmark (1979, 1983) painted a bleak picture of the corridor as it would have
appeared between about 18,000 and 12,000 rcbp. The ice sheet blocked drainage,
creating huge shifting lakes that did not sustain fish. The area would have been
even colder than today, when winter temperatures are severe. The landscape was
harsh, raw, and primitive . . . unpleasant and oppressive (Fladmark, 1983, p. 28).
Only south-facing uplands on mountains and foothills may have been relatively
comfortable refugia for plants and animals. In contrast, the north Pacific coast
was relatively warm and probably contained a discontinuous chain of unglaciated
islands, headlands, and uplands, which could have hosted terrestrial fauna, and
also allowed access to a diverse marine fauna. Fladmark (1983, p. 26) admitted,
however, that a fully glaciated Alaska Peninsula would have posed a severe im-
pediment to coastal migrants: it seems unlikely that even boat-using people could
have made it around the approximately 800-km gap between the Beringian coast
and the Kodiak refugium. If they somehow did get past this obstacle to the Cook
InletPrince Willliam Sound area, they could have kept on hopping southward
from one coastal refugium to the next (by boat); theoretically even primitive
boats could traverse the entire Pacific Coast of North and South America in less
than 1015 years (Fladmark, 1983, p. 41).
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

58 Fiedel

Mandryk (1998) suggests that the ice barrier to coastal migration would have
been present between 23,000 and 16,000 rcbp, but coastal ice receded from most ar-
eas of British Columbia and southeastern Alaska between 16,000 and 14,000 rcbp.
Bears were living on Prince of Wales Island in the Alexander Archipelago by
12,300 rcbp. If animals could support themselves in this region, so too, theoreti-
cally, could people (Fedje et al., 1996; Mandryk, 1998). Josenhans et al. (1997)
suggest, on the basis of paleobotanical data, that the Northwest Coast was suitable
for human habitation by 13,000 years B.P. (they seem to mean uncalibrated years,
or about 16,000 B.P.).
Gruhn (1988) argued for a coastal migration route, largely on the basis of
the clustering of isolated languages along the Pacific coast. However, as Meltzer
(1995) has observed, any number of postmigration developments could be respon-
sible for this late 19th-century distribution. Gruhn (1994) admitted that the coastal
theory might be untestable due to submergence of the Pleistocene coastline, but
she offered a list (1994, p. 253) of areas where hypothetical mid (ca. 50,000
30,000 rcbp) or late Wisconsin coastal migrants might have left a terrestrial
record of their passage: southwestern Washington, coastal Oregon, and California.
As Meighan (1983) and Dillehay and Meltzer (1991, p. 291) observe, even if
the littoral-focused sites of such people are now underwater, their subsistence-
settlement system probably also had an interior aspect, which should be archaeo-
logically recoverable.
So, what is the present state of the early archaeological record in these ar-
eas (Fig. 1)? The earliest known lithic industries on the North Pacific coast are
no older than 9800 rcbp (about 11,200 B.P., or 2,000 years later than Clovis).
Blades found at Anangula in the Aleutians date from 8500 rcbp. Lithic assem-
blages in the Alexander Archipelago in southeast Alaska date to 95009000 rcbp
(Ackerman, 1992). In Gwaii Haanas, off the British Columbia coast, the earliest
microblades and bifaces date to about 9400 rcbp (Fedje et al., 1996). Human re-
mains found in a cave on Prince of Wales Island date to ca. 9800 rcbp (Josenhans
et al., 1997). At Namu, on the central British Columbia coast, an assemblage
of unifacial pebble tools, leaf-shaped, unstemmed bifaces, scrapers, and gravers
dates from 9720 140 rcbp. In western Washington, a mastodon at the Manis site,
near Sequim on the Olympic Peninsula, appears to have been wounded by a bone
point at some time between 12,100 310 and 11,000 150 rcbp (i.e., Clovis-age)
(Gustafson et al., 1979; Petersen et al., 1983). A Clovis point was found near
Olympia, and others are known from eastern Washington and Oregon (Carlson,
1983). In western Oregon, the earliest site near the coast (50 km inland) dates to
barely 9000 rcbp.
California has been the scene of numerous claims of great antiquity for both
alleged stone tools and for human skeletal remains. Fractured stones at Calico
Hills, purported to be tools some 150,000 years old, and Texas Street, argued to
be more than 50,000 years old (Carter, 1980), have not convinced archaeologists
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 59

who examined the specimens (e.g., Meighan, 1983; Stanford, 1983). Skeletons
that initially yielded 14 C or amino acid racemization dates of 70,000 (Sunnyvale),
48,000 (Del Mar), 23,000 (Los Angeles), and 22,000 (Yuha) were subsequently
redated by AMS to 6300, 4900, 3560, and 16503850 rcbp, respectively (Taylor,
1991, p. 90).
In fact, there is now no unambiguous evidence of pre-Clovis people in coastal
(nor interior) California. An isolated fluted point, found in Mendocino County at a
location that would have been 10 km inland at 13,000 B.P., is the only early artifact
known from the northern coast (Erlandson, 1994). A few sites near the central coast
date from about 13,000 B.P. (11,000 rcbp), to judge from the presence of fluted
points or chipped stone crescents (found both in Clovis and Western Stemmed
assemblages in the Far West). In southern California, 21 sites and isolated finds
near the coast are dated between 10,000 and 8500 rcbp; the only one that is probably
earlier than 10,000 rcbp is an isolated fluted point from the Santa Barbara Channel
(Erlandson, 1994; Jones, 1991). Across the channel, on San Miguel Island, the
lowest clearly cultural component within Daisy Cave dates to about 10,400 rcbp.
Although a chert flake and a bone bead have been found at a still lower level, with
radiocarbon dates of ca. 15,000 rcbp on charred wood, the excavator expresses
both caution and reservations about the origin and age of these earliest materials
(Erlandson, 1998; Erlandson et al., 1996).
Just east of San Miguel (and formerly attached to it during the Pleistocene) is
Santa Rosa Island, which once seemed to offer solid evidence of an early human
presence on the coast. Radiocarbon ages in excess of 40,000 years were obtained
for ostensibly burned areas, interpreted as man-made hearths, in association with
bones of pygmy mammoths (Orr, 1968). Stone tools also were found in apparent
association with the dated material. More recent dates of about 30,000, 17,000,
and 11,800 800 also were reported for contexts with mammoth remains; a suspi-
ciously late date of 8000 250 rcbp was obtained on a sample of mammoth bone.
Skeptics have pointed out that the reported artifactual associations have not been
adequately documented, that all the dated materials seem to be redeposited, and
there is no proof of human creation of the burned areas (Erlandson, 1994; Stanford,
1983, p. 71; Waters, 1985, p. 132). Wildfires or groundwater effects may account
for the reddened soil and darkened bones (Cushing et al., 1986). In 1994, the first
nearly intact skeleton of a pygmy mammoth was found on the island (Agenbroad
et al., 1995). It is a very aged individual, who seems to have died naturally and
was not butchered. A radiocarbon date of about 12,840 rcbp has been reported.
A renewed effort to date other pygmy mammoths might yield interesting results,
as a test of the coastal hypothesis. Like canaries in a coal mine, these pony-sized
proboscideans would have been harbingers of the disaster awaiting their full-sized
cousins on the mainland, once humans arrived (Johnson, 1983, p. 513; Meighan,
1983, p. 450). In view of the seemingly inevitable rapid extinction of island fauna
when humans first appeared (e.g., moas in New Zealand and pygmy hippopotamus
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

60 Fiedel

on Cyprus), we can be pretty sure that the Santa Rosa mammoths would not have
persisted more than a few decades after the arrival of coastal hunter-gatherers. The
age of the last pygmy mammoth should tell us when humans first occupied the
coast of southern California.
The earliest coastal sites in the Americas have been found recently in southern
Peru (Fig. 2). Dates of 10,53010,770 rcbp (ca. 12,10012,900 B.P.) were obtained
for a hearth at Quebrada Tacahuay (Keefer et al., 1998). Nearby Quebrada Jaguay
280 is of similar antiquity; the lowest level yielded dates of 11,088 120 and
11,105 260 rcbp, and the overlying layer dates to ca. 10,700 rcbp (Sandweiss
et al., 1998). Both sites show intensive use of marine resources: drumfish and clams
at Quebrada Jaguay, cormorants and anchovies at Quebrada Tacahuay. Oddly, these
sites are 7 km from the probable shoreline of this period, meaning that the catch
from the sea must have been hauled to the campsites. The presence of obsidian at
Quebrada Jaguay from an Andean source about 130 km inland suggests that the
inhabitants either moved up into the Andes as part of a seasonal round or traded
with the inhabitants of the highlands. These sites demonstrate a surprisingly early
maritime orientation (assuming the charcoal-based dates are correctold wood
effects are possible in this desiccated environment), but they do not in themselves
verify a coastal migration hypothesis, as they are no earlier than fishtail sites
associated with terrestrial adaptations.

THE OVERKILL HYPOTHESIS

Mammoth, mastodon, and other American megafauna made it through the


warm climate of the last interglacial. Then, during the Wisconsin glaciation, after
ca. 70,000 B.P., temperatures rose and fell sharply many times, for periods of up to
ca. 2000 years (18 warm DansgaardOeschger interstadials and six cold Heinrich
events); but the megafauna survived these excursions. On Wrangel Island, north of
Siberia, dwarf mammoths seem to have persisted until 3700 rcbp (Vartanyan et al.,
1995). The warmcold oscillations that marked the final stages of the Pleistocene
between ca. 17,000 and 11,000 B.P. were not markedly more severe than the earlier
temperature spikes. The only obvious difference is the appearance of humans at
13,500 B.P. Paul S. Martin (1973) theorized that Paleoindian hunters, radiating
rapidly through the continent in a sort of blitzkrieg movement, slaughtered the
unwitting megafauna in a few centuries. Martins critics have wondered why only
a few species (mammoth, mastodon, and possibly horse and camel), among the
more than 30 genera that went extinct, are actually represented by skeletal remains
at Paleoindian kill or campsites. On the other hand, Paleoindians clearly did hunt
bison and caribou, both of which survived the extinction episode. Critics of the
overkill hypothesis also note that a number of bird species died out, which would be
difficult to attribute directly to human predation (Grayson, 1991). It also has been
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 61

argued that the Late Pleistocene extinction might have been a nonevent, that the
die-off was spread out over millennia. However, recent radiocarbon dates reaffirm
the suddenness of the extinction event for all species at 11,000 rcbp (13,000 B.P.), at
the onset of the Younger Dryas (Graham, 1998). In numerous exposed stratigraphic
profiles in the Southwest, a thin aquoll or black mat (spring-laid soil) directly
overlies the bones of the last mammoths. Both radiocarbon dates and climatic
correlation indicate that the black mat deposits represent a pluvial period at the
beginning of the Younger Dryas cold period, about 13,000 B.P. (Haynes, 1998;
Quade et al., 1998).
Paleontologists have often tried to explain the extinction by contrasting Late
Pleistocene and Early Holocene environments (e.g., Graham, 1998); the reduced
patchiness of the latter is thought to have posed difficulties for megafauna. It also
has been suggested that the more extreme (less equable) seasonal variation in the
Holocene might have disrupted the reproductive cycles of large mammals. The
obvious problem with such models is that, in North America, the megafauna did
not make it into the Holocene; their extinction was complete some 1500 years
earlier, at the onset of the Younger Dryas. In southern South America, deglacial
warming began about 17,000 B.P. (14,000 rcbp), and the environment seems to
have approached essentially modern conditions by 12,500 rcbp (14,500 B.P.).
Yet the inhabitants of Monte Verde butchered at least a half-dozen mastodon-like
gomphotheres around 14,000 B.P., and Paleoindian hunters were killing horses
in Patagonia at 12,800 B.P. Thus it seems that the South American megafauna
might have survived regional warming and environmental change had they not
suffered predation pressure from humans. The overkill theory has been strength-
ened by newly reported evidence that, in an analogous case, the arrival of humans
in Australia ca. 50,000 B.P. was soon followed by the extinction of giant birds and
marsupial megafauna (Flannery, 1999).
Other than human predation (combined with environmental stress caused by
climate oscillations), the only other plausible explanation of rapid extinction is the
spread of some new lethal virus or bacteria (New York Times, 1997; Hall, 1999a,
citing work of Ross MacPhee). The vector of this new disease agent would have to
be a mammalian species intruding into the Americas for the first time at the end of
the Pleistocene. Thus the likely suspects are humans or their dogs. What disease
could cause such transspecific devastation over such a huge area, in so short a time
(rabies and canine distemper have been suggested)? How would it have been passed
from humans to the indigenous animals? These questions remain unanswered. It
seems unlikely, but not impossible, that genetic analysis of megafauna bones will
allow identification of this hypothetical killer. For now, extinction due to predation
seems more likely.
An interesting variation of the overkill theory has been proposed recently
by Elin Whitney-Smith (1998). She suggests that Paleoindians would initially
have concentrated on killing large carnivores, which not only posed a direct threat
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

62 Fiedel

to them, but also were competitive predators. Once carnivore populations were
reduced only slightly (less than 2%) by human hunting, prey populations would
have grown unchecked, leading first to drastic vegetation changes and then to
catastrophic collapse when the megafauna exceeded carrying capacity. Human
predation on the remaining, drastically reduced herds could then easily result in
extinction. As there is no evidence of butchered carnivores at any known Clovis
sites, this model is likely to remain entirely speculative.

SKELETAL AND GENETIC EVIDENCE FOR


NATIVE AMERICAN ANCESTRY

The Mongoloid appearance of Native Americans is consistent with their


presumed Asian origin. Similarities include coarse black or dark brown head hair,
general sparseness of facial and body hair (although full beards were common in
some California and Great Basin groups), brown eyes, some occurrence of epi-
canthic skin folds over the eyes, and high cheekbones. The teeth of both living
and skeletal Amerinds are typified by shovel-shaped incisors and other crown and
root peculiarities that also are typical of North Asian populations; Christy Turner
(1983, 1987, 1994) has referred to this dental pattern as Sinodonty. However, Native
Americans noses are often more prominent than those of contemporary Asian peo-
ples, and many individuals lack the typical Mongoloid eyefolds. These differences
have been explained in one of two ways. W. W. Howells (1967) suggested that
Amerind ancestors emigrated from Northeast Asia at a time prior to the emergence
of the classic, specialized Mongoloids in China and peripheral regions. The earlier
Homo sapiens populations of Northeast Asia were generalized Mongoloids, with
highly variable features that sometimes resembled those of modern Caucasoids
(people of Europe and North Africa). Apart from the Americas, the more gener-
alized Mongoloids also survived on the peripheries of Asia, in Polynesia and in
Japan (the Ainu). Howells noted that the cranial traits of the poorly dated Late
Pleistocene individuals from the Upper Cave at Choukoutien (or Zhoukoudian),
near Beijing, were unspecialized (see also Kamminga and Wright, 1988), so the
classic Mongoloids were not yet dominant in northern China ca. 10,000 years ago
[but Turner (1983) notes that the Upper Cave teeth appear to be Sinodont]. This
model is consistent with a single migration event (but allows for a much later
migration of the more specialized-looking Eskimos). Alternatively, Amerind pop-
ulations have been interpreted as hybrids resulting from multiple migration waves.
The earlier migrations are speculated to have been composed of Proto-Caucasoid,
Australoid, or southern Asian Sundadont Mongoloid physical types, while subse-
quent immigrants were specialized, Sinodont northern Mongoloids. Support for
this model has been found in several facts: some of the earliest known American
crania are long and rugged, unlike the round, smooth-browed skulls of modern
Mongoloids (Steele and Powell, 1992, 1994); some marginal populations, such
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 63

as the natives of Tierra del Fuego, have peculiar, archaic-looking traits, as would
be predicted if genetic swamping of the original population was less complete
in isolated peripheral areas (Lahr, 1995); and the hairy, round-eyed Ainu can be
viewed as a distinctive Caucasoid or Australoid racial isolate in Northeast Asia,
rather than unspecialized Mongoloids.
Actual physical remains of Paleoindians are extremely rare, but a better sam-
ple is available from the Early Holocene (ca. 10,0007000 rcbp) (Table I). Unfor-
tunately, the study of the few known early crania and skeletons discovered in North
America has been severely impeded by heavy-handed bureaucratic implementa-
tion of recently enacted repatriation laws (the Native American Graves Protection
and Repatriation Act of 1990, or NAGPRA). For example, the skeleton of Buhl
woman was handed over to the Bannock Shoshoni after a brief study period (Green
et al., 1998), and only legal action by concerned scientists has so far prevented a
similar disposition of the Kennewick skeleton.
In Table I, analysts comments on Sinodonty or other Asian-like traits have
been emphasized, as a counterpoint to recent widely publicized suggestions that
Kennewick Man demonstrates an early Caucasoid presence in North America
(Morell, 1998; Rensberger, 1997). In fact, Kennewick is not uniquely early, and his
dentition has been described in preliminary assessments as Sundadont (Chatters,
1997); this would point to a southern Asian, perhaps Ainu-related, but not European
origin. Sinodonty and other Mongoloid traits have been observed in both contem-
poraneous and much older skeletons.
The American Early Holocene skeletal record, meager as it is, is still better
than the contemporaneous Asian evidence. Specimens of Chinese Homo erectus
and Neanderthal-like archaic sapiens date between 800,000 and 100,000 B.P.
Some anthropologists have cited traits shared by these fossils and recent
Mongoloids as evidence of regional continuity, but the ancestral relationship of
the pre-sapiens Asians to anatomically modern humans is debatable and seems
increasingly unlikely as genetic evidence points to an African origin and late
(post-90,000 B.P.) dispersal of Homo sapiens. The earliest known modern sapiens
specimens in northern China are still the three individuals from Zhoukoudian.
They are probably not much older than 10,000 rcbp, yet they do not display
the typical features of modern Mongoloid peoples. It is only around 7000 rcbp
that Chinese Neolithic skeletons appear indisputably Mongoloid (Brown, 1999;
Kamminga and Wright, 1988). The earliest human remains from Japan are four
skeletons from Minatogawa on Okinawa, dated to about 18,00016,000 rcbp.
Apart from the forward position of the cheekbones, the Minatogawa crania are
not very similar to the skulls of Neolithic or recent Mongoloid East Asians. Their
dentition is Sundadont, not Sinodont (Turner, 1987). The implication of these
Asian specimens is that, since the ancestors of Paleoindians probably had left Asia
prior to 10,000 rcbp (11,000 B.P.), we should not expect the earliest people in
the Americas to closely resemble the recent Mongoloids of northern Asia (Lahr,
1995).
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

64 Fiedel

Table I. Early Skeletal Remains


Site Date Description /comments Reference(s)

(A) 10,000 rcbp (11,200 B.P.) or earlier


North America
Fishbone 10,900 300; Dates on twined bark wrapped Steele and Powell,
Cave, 11,200 250 rcbp around burial; postcranial 1994; Willig,
Nevada fragments 1996
Buhl, Idaho 10,675 95 rcbp, on Stemmed knife associated; Green et al., 1998,
collagen craniofacial features fall p. 446
within the range of
American Indian or East
Asian populations
Midland, Ca. 10,000 Extremely long, high skull, Howells, 1967,
Texas 11,000 rcbp? small teeth and jaws. p. 305
Otamid-like; For all her
antiquity there is nothing
non-Indian about her
Sinodont Turner, 1983
Marmes, 10,130 300 Cranial fragments; some Sheppard et al.,
Washington 9820 300 rcbp Mongoloid features 1987; Oakley
et al., 1975,
p. 39
Anzick, Ca. 10,900 rcbp Cranial fragments; amino Stafford, 1994
Montana acid dates: 10,240 120,
10,820 100, 10,710
100, 10,940 90,
10,370 130
Wilson- 10,0009800 rcbp Steele and Powell,
Leonard, 1994
Texas
Horn Shelter, 9980 370 to Shoveled incisors; There Young et al., 1987,
Texas 9500 200 rcbp were no significant p. 295
differences between the
Horn Shelter and other
[later] central Texas
samples on cranial, post-
cranial or discrete traits
Warm Mineral 10,260 290 rcbp Cockrell and
Springs, Murphy, 1978
Florida
Mostin, 10,5007800 rcbp? 10,470 490 rcbp, fragments Taylor et al., 1985
California 10,260 340 on Possible carbonate Erlandson,
collagen, 7700 contamination, real age may 1994, pp. 270
90 on charcoal be 63003500 rcbp 271
Witt site, 11,380 70 B.P. (U-Th age, =10,000 rcbp) Willig, 1991
Tulare
Lake,
California
(Continued )
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 65

Table I. (Continued )
Site Date Description /comments Reference(s)

Arlington 10,000 310 rcbp Femora Steele and


Springs, 10,000 200 rcbp Powell,
Santa Rosa on charcoal, 1994;
Island, 10,080 810 rcbp Erlandson,
California on collagen 1994
Central and South America
Pampa de 10,200 180 rcbp, Adult and adolescent Chauchat, 1988
Fosiles 13, charcoal date
Peru
Lagoa Santa, Ca. 12,000 Sinodont Turner, 1983
Brazil 8,000 rcbp?
Lapa 10,200 220 to Cranium and postcranial Gruhn, 1991;
Vermelha 11,600 500 rcbp elements, adult female Schmitz, 1987;
IV, Brazil Prous and
Fogaca, 1998
Sueva 1, 10,090 rcbp Correal U. and
Colombia van der
Hammen,
1979
Tepexpan, 1950 rcbp (AMS date) Formerlybelieved to be ca. Taylor, 1992
Mexico 10,00011,000 rcbp

(B) 10,0007000 rcbp (11,2008000 B.P.)


North America
Sulphur 820010,000 rcbp Sinodont Waters, 1986
Springs
Woman,
Whitewater
Draw,
Arizona
La Brea, 9000 80 rcbp Skeleton; may be much younger Steele and
California Powell, 1994
Windover, 7290 120 rcbp mtDNA haplotype X found in Hauswirth et al.,
Florida, brain 1994; Purdy,
1991
Warm Mineral 75807140 rcbp Long-headed, shoveling of Purdy, 1991,
Springs, maxillary incisors p. 200
Florida,
Renier, Ca. 90008000 rcbp Cremation Mason and Irwin,
Wisconsin 1960
Eva, Stratum 4 7200 500 rcbp Minimum age on antler; Lewis and Lewis,
5 burials in bottom of 1961
stratum IV; 3 measurable
(male, 3035 yr; male,
60+ yr; female, 2530 yr);
continuity with later Archaic
inhabitants of the site
(Continued )
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

66 Fiedel

Table I. (Continued )
Site Date Description /comments Reference(s)

Kennewick, 8410 60 rcbp = Sundadont dentition, some Chatters, 1997;


Washington 7880 160 Caucasoid-like traits Taylor et al.,
reservoir-corrected, 1998
83409200 B.P.
Koster 8500 rcbp 4 adults, 3 infants Struever and
Horizon 11, Holton, 1979
Illinois
LAnse 7530 140 rcbp 12-year-old Tuck and
Amour, McGhee, 1976
Labrador
On-Your- 9730 60 rcbp; Mandible, vertebrae, pelvis Fifield, 1996;
Knees 9880 50 rcbp Josenhans
Cave, et al., 1997
Prince of
Wales
Island,
Alaska
Gordon 9700 250 rcbp on Female, mesocranic Breternitz et al.,
Creek, collagen 1971; Oakley
Colorado et al., 1975
Pelican 7900 rcbp Adolescent female, mesocranic, Wormington,
Rapids, shoveled incisors [Sinodont]; 1957 [Turner,
Minnesota conch shell gorget and antler 1983]
knife
Browns 8700 110 rcbp Dolichocranic male, red Steele and
Valley, ocher, broad Plano-like points Powell, 1992,
Minnesota 1994;
Wormington,
1957
Sauk Valley, No date Steele and
Minnesota Powell, 1992,
1994
J. C. Putnam, No date Steele and
Texas Powell, 1992,
1994
Spirit Cave, 9415 25 rcbp Mummy Jantz and Owsley,
Nevada (weighted average) 1997
Wizards 9200 rcbp mtDNA is similar to modern Jantz and Owsley,
Beach, Native Americans 1997
Nevada
Hourglass 8000 rcbp (8170 100, mtDNA comparable to Nootka, Hall, 1997
Cave, 7944 84, Maya, Yakima, Yanomamo
Colorado 7714 77 rcbp)
Gore Creek, 8250 115 rcbp No cranium Carlson, 1983,
British p. 82
Columbia
(Continued )
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 67

Table I. (Continued )
Site Date Description /comments Reference(s)

Central and South America


El Riego, 90007000 Anderson, 1967
Mexico
Arroyo Seco, 7800 115 rcbp, Politis, 1997
Argentina 7615 90 rcbp
Intihuasi, 8060 100 rcbp, On bone; 6 individuals Oakley et al.,
Argentina 7970 100 rcbp 1975
Acha, Chile Ca. 9000 rcbp Naturally mummified Arriaza, 1995
Camarones, 7000 rcbp Chinchorro mummified child Arriaza, 1995
Chile
Santo 7740 85 rcbp Cephalic index 80.57 Beynon and
Domingo Siegel, 1981
Tomb 2,
Peru
Santo 8830 190 rcbp Cranial deformation Beynon and
Domingo Siegel, 1981
Tomb 1,
Peru
Tres Ventanas 8030 130 rcbp Beynon and
Tomb 1 Siegel, 1981
Quiqche Cave 9940 200 rcbp for Some degree of shoveling Beynon and
Tomb 1 underlying level Siegel, 1981
Tequendama, Ca. 7200 rcbp Skeleton; cremations pre-8500 Correal U. and
Colombia van der
Hammen,
1977
Serra do Cipo, Ca. 9000 rcbp Schmitz, 1987
Brazil
Santana do 9460 110 rcbp Burial XII; others (men, Prous, 1992;
Riacho, women, children) mainly Prous and
Brazil 85008200 rcbp Fogaca, 1999
Toca da Janela 9670 140 rcbp Gracile woman Weber, 1993
da Barra do
Antoniao,
Brazil
Las Vegas, 82506600 rcbp 192 individuals Stothert, 1985
Ecuador

In the past few years, analyses of mitochondrial DNA (mtDNA), both sampled
from living individuals and extracted from ancient skeletons, have revolutionized
the study of the biological relationships and ancestry of Native Americans. Re-
sults obtained to date indicate the existence of five American mtDNA haplotypes,
labeled A, B, C, D, and X, each representing a female descent lineage (mtDNA
is only inherited from ones mother). All of these haplotypes are found through-
out North, Central, and South America (although X is quite rare). Each sampled
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

68 Fiedel

population seems to contain at least two, and usually three, of the four haplotypes.
However, the relative percentages of the haplotypes vary between populations, a
fact that may allow their relationships to be assessed. For example, Inuit, Haida,
and Athapaskan populations have mainly A, with minor percentages of D or C; in
contrast, B predominates in most Southwestern populations, with low frequencies
of C and D, while the A haplotype is rare or absent (Lorenz and Smith, 1996). The
ubiquitous occurrence of the five haplotypes indicates either remarkably effective
mixture of five separate genetic lineages after their arrival in the Americas or,
much more probably, a single migration event by a group initially containing all
these Asian-derived haplotypes (Kolman et al., 1996; Merriwether et al., 1994).
Several recent analyses of Y-chromosome structure (Bianchi et al., 1997, 1998;
Santos et al., 1999; Underhill et al., 1996) have indicated that all the male descent
lineages in the Americas similarly converge toward a single ancestral population.
Although the A, B, C, and D haplotypes occur widely in the modern pop-
ulations of Mongolia, central China, Tibet, and Taiwan, the highest frequencies
yet found in any Asian population have been reported recently for the Tuva and
Buryat of southern Siberia, where 72.2 and 52.4%, respectively, of sampled indi-
viduals belong to one of these four haplotypes (as compared to 48% of Mongolians)
(Derevianko et al., 1998). The B haplotype seems to be absent among the modern
people of northeastern Siberia, indicating that a population replacement proba-
bly occurred there, millennia after the departure of Paleoindian ancestors (Schurr
et al., 1999). Haplotype X, present in North American native groups including
Ojibwa, Nuu-Chah-Nulth, Sioux, and Yakima, has not yet been seen in any East
Asian population, but a distantly related mtDNA type has been reported recently
in European populations (e.g., Finns, Italians, and Druze). As haplotype X is not
accompanied by other typical European haplotypes in Native Americans, it seems
that it cannot be ascribed to early historic admixture with Europeans (Brown et al.,
1998). Furthermore, this haplotype also has been found in some prehistoric North
American skeletal populations. Perhaps, the presence of X in Europe is a legacy
of the various Turko-Mongol invasions (e.g., Huns in Italy, Mongols in Austria)
or the earlier migrations of Uralic peoples. Alternatively, it may be a trace of pos-
sible ancient genetic exchange and cultural contact between Trans-Baikal Upper
Paleolithic Caucasoids and North Chinese/Mongolian proto-Mongoloids.
Such an exchange also could explain the most recent evidence derived from
a studies of Y-chromosome DNA polymorphisms (Karafet et al., 1999; Santos
et al., 1999). The most common Native American Y haplotype, designated as
haplotype 31, is present in almost 90% of the tested males in some groups (in-
cluding populations of so-called Amerinds, Na-Dene or Athapaskan speakers, and
Eskimos). Second in prevalence is haplotype 10, found in 30% of native North
American males, and also present, at a low frequency, in Mongolia and India.
Haplotype 10 seems to be the immediate ancestor of haplotype 31, and also is
ancestral to haplotype 20, found in central Siberians (e.g., 70% of Kets and 17.4%
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 69

of Altaians) and (very rarely) in Native Americans. Haplotype 10 also appears


to have given rise to haplotype 1, which is common in European Caucasoids as
well as (subcontinental) Indians. All of these Y-chromosome variants are absent
in Chinese and Japanese males. Curiously, the Buryat, who show such a high
frequency of the ancestral American mtDNA types, do not belong to the same
lineage as Native American males. So, southern/central Siberia seems to be the
area where two ancestral gene pools overlap: one encompassing the female lin-
eages of America, northern China, and Mongolia, and the other a male lineage
that is ultimately related to the ancestors of modern Europeans, and not to East
Asian Mongoloids. Such lack of congruence between male and female ancestry
has been observed in numerous other cases and may reflect exogamous movement
of women between patrilineal bands (Karafet et al., 1999; Seielstad et al., 1998).
By reference to supposedly constant mutation rates, dates of between 41,000
and 11,000 years have been calculated for the initial emergence of the American
female lineages (Table II). Apart from these mtDNA-based estimates, a date of ca.
22,770 B.P. (with an error range from 13,500 to 58,700 B.P.) has been proposed for

Table II. Proposed Dates for Divergence of Native American mtDNA Lineages
Date Comments Reference

12,10013,200 B.P. Maximum 51007100 years for Shields et al., 1993


Haida, Eskimo, Athapaskans,
Chukchi
18,75037,500 B.P. or Excludes B haplogroup, which is Torroni et al., 1993
20,50041,000 B.P. younger
22,00029,000 B.P.; Rate based on inferred divergence Torroni et al., 1994
11,72415,456 B.P. for time of 800010,000 B.P. for
B haplogroup [mean date, Chibchan-speaking
13,590 B.P.] populations; . . . the 95%
confidence interval . . .
probably would include a date
of arrival as recent as
13,000 YBP
16,00023,000 B.P. 11,300 B.P. for Eskimo Forster et al., 1996
Athapaskan divergence
28,54941,576 B.P. One migration Bonatto and Salzano, 1997
23,00037,000 (slow rate) or Stone and Stoneking (1998, Stone and Stoneking, 1998
11,00019,000 (faster rate) p. 1163) observe that, with
respect to mtDNA,
Mongolians and Native
Americans look like members
of the same population that
began an expansion
95,000 years before the
present (48,000 years under the
faster rate)
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

70 Fiedel

appearance of the Y-chromosome haplotype 31 (Bianchi et al., 1998). However, a


date of ca. 7600 5000 B.P. has been calculated for emergence of one of the two
most common American Y haplotypes, called 1G (Karafet et al., 1999). The huge
standard errors and the differences between these various estimates underscore
the present uncertainty of molecular dating (Howell et al., 1996; Weiss, 1994).
In any case, the archaeological evidence, in my view, is most consistent with
the ca. 13,200 B.P. date suggested by Shields et al. (1993). The faster rate
estimate of Stone and Stoneking also is compatible (mean date, ca. 15,000 B.P.).
The pancontinental distribution of the B haplotype in the Americas, and its presence
in Mongolians and Siberians, effectively refutes the hypothesis that it represents a
later Pacific Rim migration. Rather, it would appear that the A, C, and D lineages
had already diverged, in East Asia, before the emergence there of the B mtDNA
haplotype. The mean age of the B haplotype, 13,590 B.P., as calculated by Torroni
et al. (1994), is remarkably close to the corrected date for the earliest Clovis sites
[as well as Turners (1986) dental divergence date of ca. 14,00013,500 B.P.].
How can one account for the degree of anatomical and genetic diversity of
recent Native Americans? A certain amount of regional isolation, over the course
of 13,500 years, could have resulted in both differences within the New World
and distinctive American traits that are not found in recent Asian Mongoloids. For
an instructive comparison, consider the distinctive features of the Polynesians, as
compared to southern Chinese or Filipinos; yet the split between these populations
probably occurred less than 6000 years ago (Bellwood, 1987) (although in this case,
genetic input from Melanesians was probably an important factor in changing the
phenotype). To some degree, founder effect, a nonrepresentative sampling of the
ancestral population due to the small size of the emigrant group, can explain differ-
ences such as the virtual absence of blood type B in Amerinds south of the Arctic.
Yet the founding Paleoindian band must already have contained a fair amount of
both genetic and phenotypic diversity; in addition to at least five women, each
belonging to a distinct mtDNA lineage, this group included men who were de-
scended from Eurasian hunters and perhaps looked vaguely like Upper Paleolithic
Europeans. Isolation of far-ranging groups after several generations may have ac-
centuated preexisting heterogeneity. Some physical anthropologists contend that
they can recognize, based on cranial traits, three to five distinct populations in the
Americas at 10,000 rcbp (Morell, 1998). Georg Neumann (1952) distinguished
eight cranial types in North America: Otamid (supposedly the oldest, exemplified
by skulls from the Texas Gulf Coast), Iswanid, Ashiwid, Inuid, Deneid, Lenapid,
Walcolid, and Lakotid. It should be noted, however, that Neumanns sample of
10,000 skulls was composed of much more recent specimens than the 7000- to
10,000-year-old crania examined in more recent analyses. It also is intriguing that
linguists recognize seven macrophyla in North America: Uto-Aztecan, Hokan,
Penutian, Macro-Siouan (including Iroquoian and Caddoan), Algonquian, Na-
Dene (Athapaskan), and Eskimo-Aleut. Linguistic differentiation from a presumed
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 71

ancestral Amerind language (Greenberg, 1987; but see Goddard and Campbell,
1994) could have begun soon after the rapid expansion at 13,000 B.P.

SEARCHING FOR CLOVIS ANCESTORS: THE ARCHAEOLOGY


OF LATE PLEISTOCENE NORTHEAST ASIA

The genetic profile and the skeletal and dental morphology of Native
Americans thus clearly indicate their Asian ancestry. What does the archaeol-
ogy of Northeast Asia, the obvious staging area for Paleoindian migration, reveal
about the possible date of migration and the likely cultural baggage of the people
who crossed the Bering land bridge?
If recent assertions of great antiquity for Chinese sites are valid, and the
reassessment of the age of Homo erectus in Java is correct, it would seem that
hominids were already present in East Asia by 1.8 million years ago. A more
conservative estimate would be about 1 million years. Homo erectus was certainly
living in northern China, near Beijing (Peking), by 300,000 years ago, and may
have ranged much farther north. Possible pebble tools have been excavated at
Diring Yuriakh in the Lena River Basin of Siberia. Claims of great antiquity by
the Russian archaeologist, Yuri Mochanov, have been regarded skeptically, but
thermoluminescence dates now suggest an age between 260,000 and 370,000 B.P.
for this site, at 61 N latitude (Waters et al., 1997). If Middle Pleistocene hominids
were capable of survival so far north, what would have prevented them from
crossing Beringia and entering the Americas at a comparably early date (Butzer,
1991, p. 140)?
Perhaps Homo erectus did get into the New World, but there is no acceptable
archaeological evidence of a mid-Pleistocene human presence. However, Bryan
(1978) photographed and described a poorly provenienced H. erectus-like cra-
nium that subsequently disappeared from the Brazilian museum, where it had
been stored. Also in Brazil, a few alleged pebble tools have been reported from
Level IV of Esperanca Cave, associated with fossil bones dated to ca. 295,000 B.P.
by the uraniumthorium method (de Lumley et al., 1988; see Lynch, 1990, for
a skeptical assessment). If 250,000-year-old human occupation sites exist in the
Americas, they should be as recognizable and recoverable as Old World sites of
this age. Noting that there has been little controversy about Australian sites with
relatively crude toolkits dating earlier than 30,000 years ago, Beaton (1991, p. 213)
comments, Few, if any, American archaeologists would not recognize an industry
such as the Southeast Asian Pacitanian . . . or the Australian Core Tool and Scraper
Tradition. . . . In fact, Australian archaeologists have never truly searched for such
evidence . . . it most often came to hand via the usual sourcesgeologists, farm-
ers, amateur collectors, and otherswho, along with Australian archaeologists,
have no trouble in identifying the unifacial industry. However, if the pre-Clovis
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

72 Fiedel

occupations of Pedra Furada and/or Monte Verde are confirmed, we must ponder
the implications of Butzers (1991, p. 143) observations about the ostensible tools
at those sites: At face value, the Chilean and Brazilian lithics would rarely be
recognized if present among a surface gravel lag. Even if recognized, their human
origin would be difficult to demonstrate to the general satisfaction of lithic spe-
cialists. Indeed, the artifactual character of the quartzite tools at Pedra Furada
has been challenged (Meltzer et al., 1994), and few archaeologists accept claims
that selected pieces among the fortuitously broken gravels at Calico Hills or Texas
Street in California are artifacts of mid-Pleistocene age.
Apart from the archaeological record, an increasingly strong case is being
constructed, based mainly on various molecular genetic analyses, that Homo erec-
tus and their Neanderthal-like descendants in Asia were replaced by anatomically
modern Homo sapiens, whose migration out of East Africa may have begun about
80,00060,000 years ago (e.g., Watson et al., 1997). Although the relationship
of anatomical and technological evolution is a complicated issue, the appearance
of Upper Paleolithic toolkits, dominated by blades, generally (but not always)
marks the presence of Homo sapiens. Such blade-dominated assemblages, along
with bone tools and ornaments, abruptly replaced Levallois-Mousterian indus-
tries in southwestern Siberia about 43,00035,000 rcbp (Goebel, 1993; Goebel
and Aksenov, 1995), although the Mousterian may have persisted in the Altai
Mountains to as late as 29,000 rcbp (Kuzmin and Orlova, 1998). If 43,000 rcbp
(ca. 46,000 B.P.) is the date of entry of Homo sapiens into Northeast Asia, it repre-
sents the earliest possible baseline for Paleoindian ancestry. Remarkably, this date
(admittedly at the outer limits of radiocarbons utility) is very close to Stone and
Stonekings fast rate estimate of 48,000 B.P. for expansion of the ancestral
Mongolian-American population.
Reliable early dates from Upper Paleolithic sites near Lake Baikal and the up-
per Lena River (Fig. 3) include 37,360 2000 from Arta 2, >32,865 and 30,100
150 rcbp from Ust-Kova, 29,895 1790 from Level 2 of Varvarina Gora, and nu-
merous dates of ca. 23,00026,000 rcbp (Kuzmin and Orlova, 1998; Kuzmin and
Tankersley, 1996). The site of Malta, best known for its carved bone figures of
swans and of humans in parkalike clothing, has yielded dates of ca. 20,000 rcbp
on bone samples. About 900 km to the east, in the Aldan River area of Yakutia,
occupation of Ust Mil 2 and Ikhine 2 may have occurred as early as 30,000
35,000 rcbp, but the possibility that the dated wood from those sites was inher-
ently old renders this dating suspect (Yi and Clark, 1985). A more secure date
for initial settlement of this region may be the 18,300 180 date on wood from
Verkhne-Troitskaya; after this isolated date, the next good dates are 15,200 300,
on charcoal from Avdeikha, and 14,000 100, also on charcoal from Dyuktai
Cave, level 7b. Goebel (1998) suggests that at the glacial maximum, ca. 18,000 rcbp
(22,000 B.P.), Siberia became so cold that it was virtually uninhabitable [the Guliya
ice core from Tibet indicates that central Asia became sharply colder at 28,000 B.P.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 73

(ca. 24,000 rcbp), and this core registered the coldest century since 70,000 at
ca. 18,500 B.P. (ca. 15,000 rcbp)] (Thompson et al., 1997).
Just east of the Aldan sites, the Verkhoyansk Mountains seem to mark an eco-
logical boundary (Hoffecker et al., 1993), which apparently was not breached by
western Siberian people until the sudden global warming at 12,500 rcbp
(14,700 B.P.). Only two sites tentatively dated earlier than 11,000 rcbp are known
from this far northeastern section of Siberia (or western Beringia): Berelekh and
Bolshoi Elgakhchan. At Berelekh, small blade fragments were found stratified
above a concentration of mammoth bones; the bone bed produced dates of about
13,000 rcbp. However, the fauna associated with the artifacts were of modern
species, except for some eroded-looking bones that were probably scavenged from
old deposits. A broken ovate biface and a microcore were found on the riverbank.
The occupation of Berelekh probably is later than 12,000 rcbp (Mochanov and
Fedoseeva, 1996b). Bolshoi Elgakhchan, on the Omolon River, consists of two
components; the older, 1, includes large blades, scrapers, bifaces, and projec-
tile points (Kiryak, 1996). It has been compared both to the Nenana assemblage
from Walker Road, Alaska (Hoffecker et al., 1993), and to artifacts from Ushki I
layer 7; these comparisons suggest an approximate age of ca. 14,00011,000 rcbp
for Bolshoi Elgakhchan 1.
The early Siberian Upper Paleolithic toolkits were dominated by macroblades,
but it appears that, probably as early as 24,000 rcbp (Kuzmin and Orlova, 1998)
and certainly by ca. 18,000 rcbp (Goebel et al., 1991), microblade-dominated as-
semblages of the Dyuktai culture were prevalent all over Northeast Asia. Dyuktai
assemblages also typically include well-made bifaces. The latter are credible proto-
types for Paleoindian bifaces, although the Asian points were not fluted. A crudely
fluted point has recently been reported from the site of Uptar in Siberia, but it is
a unique specimen in a context that may not be much older than 8300 rcbp (King
and Slobodin, 1996). Another major difference between Dyuktai and Paleoindian
assemblages is the absence from the latter of microblades and the small boat- or
wedge-shaped cores from which they were struck. Most archaeologists view this
distinction as so significant that it would appear to preclude Dyuktai as a candidate
for Paleoindian ancestry. In contrast, the Dyuktai roots of the Paleoarctic tradition
or Denali complex of Alaska (eastern Beringia) are universally acknowledged.
Paleoarctic lithic assemblages contain the same kinds of microblades and cores
as Dyuktai. F. H. West (1981, 1996) has been the foremost proponent of the view
that the Paleoarctic tradition gave rise to Clovis (see also Carlson, 1991; Morlan,
1991). He is compelled to admit that the complete abandonment of microblade
technology by Clovis is a problem, but suggests that a sort of cultural mutation
must have occurred, perhaps during the passage through the ice-free corridor.
Critics of Wests model have pointed out that the Paleoarctic tradition is
too late to be ancestral to Clovis, because it dates after 10,700 rcbp in Alaska
(Hoffecker et al., 1993, p. 51). They identify the slightly earlier Nenana complex
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

74 Fiedel

of central Alaska as a more credible Paleoindian ancestor because (1) it may be


marginally earlier than Clovis, and (2) it contains various tools on blades and flakes,
not microblades. However, there is still the small problem of fluted points. Nenana
assemblages do include small triangular and teardrop-shaped points, but these are
a far cry from classic Clovis fluted points. It is even possible that Nenana is only
a contemporaneous functional variant of Denali and not a separate culture at all.
This possibility is raised by recently reported dates of 11,660 rcbp for a Nenana-
like component with microblades at Swan Point. On the other hand, a Nenana
assemblage with no microblades (dated 11,120 85 rcbp) is stratified below a
Denali component (10,690 250 rcbp) at Dry Creek (Goebel et al., 1991).
It is important to note that there is no evidence of mammoth hunting at any of
the early Alaskan sites. Mammoth ivory is present, both at Broken Mammoth and
Swan Point, but in each case radiocarbon dates have shown the ivory to be cen-
turies earlier than the other material at the site (15,800 rcbp at Broken Mammoth,
12,060 rcbp at Swan Point). The inhabitants seem to have been scavenging the
tusks of long-dead mammoths as raw material for tools. They were actually hunt-
ing elk, bison, sheep, and smaller game, particularly waterfowl, which account for
about 60% of the meat (Yesner, 1998b). In western Beringia, the youngest dated
mammoth is a specimen from Yuribey, at 10,000 70 rcbp (Ukraintseva et al.,
1996). But perhaps the eastern mammoth died off more rapidly, so that none were
available to Nenana hunters after 13,900 B.P. This allows us to hypothesize that
the invention of the Clovis point and its related delivery system (bone foreshafts,
etc.) resulted from the initial encounter of Nenana or Denali people with living
mammoth, south of the ice sheets (Pearson, 1997b).
If we assume that Clovis did not mutate from the DyuktaiPaleoarctic mi-
croblade tradition, where can we find other credibly ancestral lithic traditions in
Asia? Several authors have looked westward to the Russian Upper Paleolithic.
Kostenki, Sungir, Biryuchka, and other sites of the Streletskayan complex, dated
ca. 35,00024,000 rcbp (Bradley et al., 1995), contain bifacially thinned trian-
gular points and retouched blades that resemble both Nenana and Clovis tools
(Bradley, 1997; Haynes, 1982; Pearson, 1997b). However, these sites, north of
the Black Sea, are thousands of kilometers west of Beringia and some 14,000 or
more years earlier than Clovis. The Malta-Afontova sites of the Lake Baikal and
Yenisei regions are eastern outliers of this Russian Upper Paleolithic tradition; at
Malta, it may have lasted beyond 15,000 rcbp. The Ust-Kova site in the northern
Angara region of central Siberia, dated to 23,920 310 rcbp, contains a blade-
dominated industry, with bifaces, that has been compared to Nenana assemblages
(Goebel et al., 1991) and also to Clovis (Tankersley et al., 1996). It must be noted,
however, that Turner (1994) has examined the teeth of skeletons from Kostenki,
Sungir, and Malta and found an absence of the distinctive Sinodont dental charac-
teristics. This would seem to eliminate the Trans-Baikal people as direct ancestors
of the Sinodont Paleoindians. Yet Turner (1983, p. 156) did observe a minute
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 75

Caucasoid quality in Paleoindian dentition, and the genetic evidence now points
very clearly to the area of Lake Baikal and the Altai as the ancestral homeland
of Native Americans (Karafet et al., 1999) (or at least the last surviving pocket
of a once more widespread Northeast Asian population, subsequently displaced
elsewhere in the region). Perhaps proto-Paleoindians emerged through genetic and
cultural exchanges around Lake Baikal, between Mongolian natives (mainly fe-
males?) and proto-Caucasoids (mainly males?) who carried the Russian Upper
Paleolithic complex into the region.
Looking even farther west, Dennis Stanford (1998) has noted similarities
between Paleoindian points and the beautifully chipped bifaces of the French
and Spanish Solutrean. However, the Solutrean complex is succeeded by the
Magdalenian around 16,500 rcbp (19,000 B.P.), some 6000 years before Clovis,
and a hypothesized ocean voyage along the rim of the North Atlantic appears very
improbable. The absence of Sinodonty in the Cro-Magnon remains from Western
Europe, as in Eastern Europe, also seems to preclude any direct role in Clovis
ancestry.
Although Y-chromosome analyses have complicated the issue, the dental
and mtDNA evidence (Kolman et al., 1996; Turner, 1994) points to Mongolia
and adjacent northern China as the Paleoindian homeland. In northern China and
Japan, there is increasing evidence of a biface and macroblade horizon immediately
preceding the expansion of microlithic industries. In Japan, leaf-shaped bifaces
were already present before 31,000 rcbp, in level 15 of Fukui Cave, in Kyushu.
They are found in central Honshu around 13,000 rcbp. Microblades appear there
around 12,000 rcbp, and they occur by ca. 12,500 rcbp in-Kyushu (Ikawa-Smith,
1978). In Fukui Cave, horizon 3, the microliths are associated with early pottery.
Dates of ca. 13,000 rcbp for pottery also have been reported from the sites of
Khummi and Gasya in the Amur Basin, in the Russian Far East (Kuzmin et al.,
1997).
As recently as 1991, Butzer (1991, p. 44) could state that foliate points
are absent in China. However, recent research in northern China has shown that
assemblages containing mainly macroblades and leaf-shaped bifaces, with some
microblades, were present at 12,700 rcbp, but that the microlithic element be-
came predominant after 11,600 rcbp (Elston et al., 1997). The earliest dates for
microlithic industries in China are problematic; some Chinese scholars (e.g., Pei,
1985) have argued for an age in excess of 20,000 years, but Elston et al. express
skepticism about such early dates. Elsewhere in Northeast Asia, however, the tran-
sition from macroblades to microblades seems to have occurred before 20,000 rcbp,
so the early Chinese dates would fit within the broader regional trend.
The emerging evidence suggests the possibility that not-yet-fully Mongoloid
makers of bifaces and macroblades in Mongolia, eastern Siberia, or North China
may have faced competitive pressure from makers of microliths and pottery (pos-
sibly Mongoloids, or perhaps Ainu-like people) between 13,000 and 12,000 rcbp
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

76 Fiedel

(about 15,00014,000 B.P.). Did some of these proto-Mongoloid biface makers


escape northeast into Beringia?
The enigmatic Ushki Lake sites of central Kamchatka may have some bear-
ing on this scenario. Layer 7, the lowest occupation level at Ushki I, contained
stemmed bifacial points, one leaf-shaped point, burins, end and side scrapers, and
stone beads. Notably absent are microblades and wedge-shaped cores, which occur
in the overlying Layer 6 (Dikov, 1996; Dikov and Titov, 1984). Dates of 14,300
13,600 rcbp (ca. 17,000 B.P.) were obtained for Layer 7, although an anomalous
date of 9750 rcbp also has been reported (Hoffecker et al., 1993). Layer 6, only
some 30 cm above Layer 7, has produced dates of 10,860 to 10,360 rcbp (12,900
11,400? B.P.). The Ushki assemblage does not look much like Clovis, and its age is
not certain, but at least it seems to show that nonmicrolithic cultures were present
in Northeast Asia as late as 17,000 B.P. Hypothetical Upper Paleolithic migrants
moving north from China might have traveled northeast through the Amur River
Basin and along the western side of the Sea of Okhotsk (Fig. 3) before arriving in
Kamchatka [this southern route was the one favored by Chard (1960)]. The site of
Ustinovka 1 has not been dated, but comparisons with other assemblages suggest
that the lower component dates to about 16,00012,000 rcbp, and the upper to
about 11,000 to 10,000 (Vasilievsky, 1996). Artifacts from both components, such
as sidescrapers, endscrapers, retouched blades, and blade cores with prepared plat-
forms, appear very similar to tools from Paleoindian assemblages. However, these
are not very specific resemblances; there are no Clovis-like points in the Ustinovka
assemblage, and at least the later complex includes microcores, which are absent
from Clovis. An early microblade industry is reported at Selemdga in the mid-
dle Amur River Basin, with initial dates of ca. 24,00020,000 rcbp (Derevianko,
1989). On the other hand, at Khukhtuy III, near the western shore of the Sea of
Okhotsk, a Dyuktai-like assemblage contains bifaces, but no microcores. There
are no radiocarbon dates for this material, but a date of ca. 11,000 rcbp, near the
end of the Dyuktai culture, has been suggested on typological grounds (Mochanov
and Fedoseeva, 1996c). It would probably be rash at this stage of research to rule
out the possibility of significant regional variation in Dyuktai assemblages. Rather
than postulate the rapid abandonment by initial Clovis people of one of the former
mainstays of their toolkit, it seems preferable to derive them from a conserva-
tive cultural variant that, even if familiar with microblade technology, had always
preferred to use macroblades when suitably large lithic sources were available.

MODELS OF PALEOINDIAN SOCIAL NETWORKS,


MIGRATION, AND COLONIZATION

Lithic artifacts represent but a small fraction of the entire cultural inventory of
the Paleoindians, but unfortunately they are almost all that is left for archaeological
study. Artifacts made of perishable materials such as leather, wood, and fiber have
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 77

vanished (except at Monte Verde). Human remains are extremely rare, as are the
bones of their prey. Archaeologists exult when they can assign any minute calcined
bone to a particular species hunted by Paleoindians (e.g., Storck and Spiess, 1994).
The spectacular associations of Clovis points with mammoth skeletons in the
southern Plains perhaps have skewed our interpretation of Paleoindian subsistence
in the direction of specialized big game hunting. No doubt, the typical diet included
numerous items obtained by less dramatic means: berries, roots, fish, etc. [as
indicated, for example, by remains from Shawnee-Minisink (McNett, 1985)]. In
the absence of better-preserved representative diet samples, we will probably never
be able to assess the relative importance of hunting and other subsistence activities.
Turning to the lithic toolkit, we can at least note the absence or paucity of grinding
stones that might be needed to process seeds. Instead, Clovis assemblages contain
points, knives, side and endscrapers, gravers, and perforatorsa toolkit manifestly
well designed for hunting, butchering, processing hides, and making tools of wood
and bone.
It is actually rather surprising that Clovis people seem to have been almost
as obsessive about their stone artifacts as archaeologists are. They were not averse
to traveling long distances of 300 km or more to obtain the best-quality cherts
and chalcedonies for their points. This fact has led researchers to some interesting
inferences. (1) Paleoindian seasonal movements were tethered to the quarries
where they obtained chert, a vitally important material (Gardner, 1974, 1989).
(2) If a Clovis assemblage at a site does not contain artifacts made of lithic material
from a high-quality source located a short distance away, this may indicate that it
represents a camp of the initial explorers in the area, who had not yet discovered
the source [e.g., Gramly and Funk (1990), suggest this about the Lamb site in
western New York]. (3) The distance from camps to lithic sources may indicate
the extent of a bands annual movements within their territory [although special
procurement expeditions and exchange with neighbors also could account for the
occurrence of exotic material (Hayden, 1982; Seeman, 1994; Spiess et al., 1998)].
Ethnographic studies of the 19th- and 20th-century Cree and Athapaskans
(Helm, 1968; Rogers and Leacock, 1981), who lived by hunting caribou, fishing,
and trapping fur-bearing mammals, provide some insight into the demographic
parameters of a boreal forest adaptation that may have been similar to that of
northeastern Paleoindians (Custer and Stewart, 1990). The typical Subarctic pop-
ulation density is about one person per 200 km2 ; at this density, it would have
taken about 125,000 Paleoindians to fill North and South America below the ice
front. Assuming doubling per generation (every 20 years), it takes 160 years for
an initial population of 500 to reach this size. In radiocarbon time, this process
would unfold within a 1 time frame; that is, it would appear instantaneous.
Both empirical observation of Australian bands (Birdsell, 1968) and com-
puter simulation (Wobst, 1976) suggest that the minimal effective size of human
mating networks (and dialect tribes) is about 500 individuals. Based on these
numbers, we can speculatively model a multitiered Paleoindian social landscape,
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

78 Fiedel

with microbands of ca. 25 people in 5200-km2 territories, aggregating seasonally


in macrobands of ca. 100 to 150. Four to six macrobands might form a network
of about 500 people for exchange of spouses and trade items. This social entity
would encompass a territory of about 100,000 km2 , envisioned as an approximate
circle with a radius of about 180 km [compare this to the archaeological record of
Terminal Pleistocene Northwest Europe, where three stylistically defined ethnic
groups each occupied an area of ca. 100,000 km2 , implying a ca. 180-km radius
(Price, 1991, p. 199)]. In the continental United States, there might have been
100 such territories, with a total population of about 50,000, around 12,500 B.P.
This social landscape would represent a late, stabilized situation; social organi-
zation during initial expansion into unbounded territory could have been quite
different.
Several case studies of lithic distributions permit testing of this model against
the actual archaeological record. In the area surrounding the Williamson quarry in
southeastern Virginia, McAvoy (1992) recognized 58 Clovis points made of chert
visually attributable to this source. Eighty percent (46 of 58) of these points were
found within an 80-km radius. This may imply a ca. 20,000-km2 territory, occu-
pied by a 100-person macroband (McAvoy suggests 75 people in a 13,000-km2
territory). Anderson (1996, p. 37) observes that similar artifact concentrations
from identifiable lithic sources tend to occur at roughly 250- to 400-kilometer
intervals. . . . This may indicate the geographic scale over which discrete Paleoin-
dian groups ranged, probably at the band level and possibly incipient macroband
level of organization. Anderson (1995) hypothesized the existence of 15 Mid-
dle Paleoindian (ca. 12,500 B.P.) macrobands east of the Mississippi, each in a
territory of ca. 20,00080,000 km2 . Haynes (1980a, p. 118) observed that the
figure of 300 km (200 mi) for the maximum distance from lithic sources turns up
repeatedly in the literature. At the Gainey site in southeastern Michigan (Simons
et al., 1984), 60% of the tools were made from Upper Mercer chert, derived from
outcrops in Ohio, about 350 km to the south. Similarly, over 65% of tools from
the Paleo Crossing site in northcentral Ohio (dated to 10,980 75 rcbp) (Brose,
1994) were made from Wyandotte and Dongala cherts, obtained about 480 km to
the south (Stothers, 1996). At the Bostrom site in southwestern Illinois, 60% of the
artifacts were made of nonlocal stone, including Attica chert from westcentral
Indiana (240 km to the northeast) and Dongola and Kaolin cherts from southern
Illinois (160 km southeast). Knife River flint from southwestern North Dakota, al-
most 1600 km away, also was recognized at Bostrom (Tankersley, 1998); one of the
concave-based points found in the Lamb cache in western New York also appears
to be made of this material (Gramly and Funk, 1990). At Bull Brook in northeast-
ern Massachusetts, the predominant material is St. Albans chert from Burlington,
Vermont, 305 km to the northwest (Spiess et al., 1998, p. 204). It is improbable that
exchanges between neighboring bands could account for such high percentages
of cherts from distant sources. Direct procurement seems more likely, but there is
no obvious way to determine if this was embedded in the normal subsistence
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 79

round, or if, instead, special-purpose task groups made periodic expeditions to the
stone quarries (Spiess et al., 1998).

How Rapid Was Initial Colonization?

Whitley and Dorn (1993) argued that, even setting aside dubious pre-Clovis
sites, the oldest unquestioned dates for South American Paleoindians required
a migration rate that was implausibly rapid for hunter-gatherers, and was even
unparalleled in the history of more complex societies. They used this theoretical
argument to bolster their claims for rock-varnish dates of ca. 15,00025,000 rcbp
for petroglyphs and lithic artifacts in the western United Statesdates that Dorn
(1996) tacitly disowned a few years later. However, their general line of reasoning
is still valid. The dates for South American Paleoindian sites are a crucial constraint
on the temporal frame of Paleoindian migration.
When Junius Bird (1938) discovered remains of the earliest hunters at Fells
Cave in Tierra del Fuego, he assumed that their fishtail (or Fell 1) points were
derived from an ancestral Clovis tradition. This inference has been supported by
subsequent finds of very similar points 5600 km to the north, at El Inga in the
highlands of Ecuador (Mayer-Oakes, 1984), and even farther north, in Panama
and Costa Rica (Ranere and Cooke, 1991; Snarskis, 1979). Several recent studies
have shown that the style and production trajectory of fishtail points are indeed
derived from North American Clovis and that the geographic ranges of Clovis
and fishtail types overlap slightly in Central America (Morrow and Morrow, 1997;
Pearson, 1998; Ranere, 1997). Fell 1 fishtail points in the Southern Cone have been
securely dated at ca. 10,30011,100 rcbp, or ca. 12,30013,000 B.P. [e.g., at Cueva
del Medio in Patagonia (Nami, 1996) and at Cerro La China and Cerro El Sombrero
in the Pampas (Flegenheimer, 1986/1987; Flegenheimer and Zarate, 1997)]. Sev-
eral ostensibly older dates of ca. 13,00012,000 rcbp also have been reported for
Los Toldos 3, Level 11 (12,600 600), Cueva del Lago Sofia (11,570 60 and
12,990 241), and Piedra Museo (ca. 12,700) (Politis, 1997), but all these dates
seem to be erratic, as associated assemblages either are clearly Fell I or are probable
functional variants that lack the diagnostic points (Nami, 1994).
The earliest relatively precise dates yet obtained for North American Clovis
are two dates of ca. 11,550 rcbp (about 13,20013,600 B.P.) for the Aubrey site in
Texas (Ferring, 1995). Most other dates with small standard deviations fall between
11,200 and 10,700 rcbp. All of these dates could fall within a century of calendrical
time at the onset of the Younger Dryas cold period, about 13,00012,900 B.P.,
but evidence of a date reversal (Fiedel, 1999) allows alternative correction to
ca. 13,200 B.P. Thus the longest possible temporal span separating the initial
inhabitants of the Southern Cone from the earliest Clovis hunters in Texas is about
1300 years (13,60012,300), but a more likely span is only 400 years (13,200
12,800).
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

80 Fiedel

To get from the southern end of the ice-free corridor to Texas by the most
direct route, Clovis ancestors would have covered about 2100 km. They could have
done this in less than 100 years. In fact, they may previously have trekked through
the ice-free corridor in a matter of months, leaving no archaeological record. At a
reasonable pace of 24 km per day, it would have taken 80 days to walk the 1900 km
through the corridor.
The ethnohistoric record of recent, territorially constrained hunter-gatherers
offers no analogy for such rapid migration. However, archaeology provides an
analogy from much later Arctic hunting cultures. Thule whale hunters, ances-
tral to recent Inuit or Eskimos, moved along the coast from northern Alaska to
Greenland (ca. 24003200 km) in less than 150 years (A.D. 9001050), and in the
process competed against and replaced people of the long-resident Dorset culture
(McGhee, 1984). If the initial Clovis colonists, setting out from Panama, moved at
a comparable rate on an essentially linear route of ca. 6400 km along the Andean
chain, they would have reached Tierra del Fuego within 300 or 400 years, as
required by the chronological evidence.
In both cases, migration seems to have been part of a response to ameliorat-
ing climate. Thule movement occurred during the Medieval Warm Period. Emi-
gration of Clovis ancestors from Beringia probably occurred around the middle
of the BllingAllerd period, a dramatic warming that began about 14,700 B.P.
(12,600 rcbp). However, maximal Clovis expansion within North America oc-
curred at a time of climatic stress, during the Intra-Allerd Cold Period (13,250
13,100 B.P.) and the ensuing 200-year warm period that created a drought in the
Southwest.
If the Monte Verde-Joboid people had begun to colonize South America
before 14,000 B.P., they were rapidly replaced, at least in the Andes and Southern
Cone, by the Clovis-descended fishtail point makers, just as Thule replaced Dorset
in the Arctic. Fishtail-like points have been found at Salto Chico, 10 km south of
Monte Verde (Dillehay, 1997).
Assuming that Paleoindians were descended from hunting peoples of north-
ern Eurasia and Beringia, they had a long tradition of primary dependence on meat.
Plant foods were simply insufficient to support human life in Arctic and Subarctic
latitudes. Furthermore, a colonizing population could rely on preexisting knowl-
edge of the behavior of large, far-ranging mammals, while it would have taken
years to acquire intimate knowledge of the distribution and uses of territorially
restricted and variable plant species. Because of rapidly changing environments,
Late Pleistocene fauna would have frequently shifted their ranges and density.
These changes would have prevented Paleoindians from settling in at a particu-
lar location and predicting the scheduled movements of their prey. Thus hunting
of megafauna and frequent movement would have been elements of a successful
adaptation for early Paleoindians (Kelly and Todd, 1988).
There has been much discussion about whether Paleoindians should be char-
acterized as specialist big game hunters or more eclectic generalized foragers.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 81

As already suggested, Arctic and Subarctic ancestral Paleoindians would have had
little choice but to be hunters, supplementing big game with smaller mammals,
fish, and berries in season. It is noteworthy that the Nenana people of central Alaska
relied heavily on the meat of waterfowl (Yesner, 1998b). Observation of migratory
waterfowl, flying south through the ice-free corridor, might even have convinced
Alaskan proto-Paleoindians that habitable land lay beyond the ice sheets; and these
birds, drawn to the vast lakes left in the wake of the melting ice, also could have
provided food during the trek south. As the Paleoindians moved southward into
temperate deciduous forests, then into equatorial montane forests and even rain-
forests, many more potential plant foods would have become accessible, while
the density of megafauna diminished. Apart from this latitudinal variation, the
collapse of megafaunal populations at 13,000 B.P. (11,000 rcbp) would have com-
pelled Paleoindians everywhere to adopt more broad-spectrum diets. Evidence
has been accumulating in South America of a rapid change in subsistence after
11,000 rcbp. Roosevelt et al. (1996) present evidence of intensive consumption
of tree fruits, along with fish, birds, mollusks, reptiles, amphibians, and small and
medium-sized mammals, in the Monte Alegre culture in the Brazilian Amazon by
10,600 rcbp (ca. 12,30012,800 B.P.) (Fiedel, 1996a) and perhaps even earlier at
ca. 11,000 rcbp. Kipnis (1998) presents dates from the basal layers of several cen-
tral Brazilian rockshelters that also indicate broad-spectrum use of fruits, roots, and
small game starting between ca. 11,000 and 10,500 rcbp. Some sites in this region
have produced even older dates, e.g., Lapa do Boquete (three dates ca. 12,000 rcbp
and two ca. 11,440 rcbp), Santana do Riacho (11,960 250 rcbp), and Sitio do
Meio (two dates ca. 12,300 rcbp). It remains to be seen whether these earlier dates
are misleading outliers, like some of the early Southern Cone dates or, instead,
are associated with a Monte Verde-like pre-fishtail foraging adaptation. It should
be noted that the earliest assemblages at Lapa do Boquete and Lapa dos Bichos
(10,450 70 rcbp) contain endscrapers and rare bifacial projectile points, while
the toolkit from Santana do Riacho includes endscrapers, sidescrapers, borers, and
rare points. Endscrapers, most effective for hideworking, seem like a holdover
here from an earlier hunting-focused toolkit. The contemporaneous appearance of
a marine adaptation in coastal Peru between 11,000 and 10,700 rcbp has already
been noted. What should be emphasized is the capability of hunting peoples to
quickly and radically alter their subsistence modes. In an analogous case, Thule
whale hunters abandoned whaling and became reliant on fishing, sealing, and cari-
bou hunting within 200 years after their initial peopling of the Canadian Arctic
(Fiedel, 1998; McGhee, 1984). Evidence of broad-spectrum economies, even as
early as 11,000 rcbp, is not incompatible with initial migration a few hundred years
earlier.
The initial immigrants, whether their route was interior or (less probably)
coastal, had trekked about 2400 km without stopping for very long. The ice-free
corridor, though passable, probably lacked sufficient resources to sustain long-
term settlement before 11,000 rcbp. Once through the corridor, there is no reason
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

82 Fiedel

to assume that Paleoindians would have changed their habits. Rather than imagine
a slow budding process across generations, with daughter bands remaining in
close proximity to the parent band, we may suspect that each generational split of
growing bands might have resulted in a movement of daughter populations many
hundreds of kilometers. This sort of long-distance movement has been termed
leap-frogging (Anderson and Faught, 1998; Anthony, 1990; Fiedel and Anthony,
1979).
Paul S. Martin envisioned human hunters advancing as a destructive wave
through the sea of indigenous American fauna. This model raises some interesting
questions. Did anyone remain behind after the wave front had passed through an
area? If local extinction had occurred, what did this remnant human population
do for subsistence? When the human wave front came up short against the edges
of the land mass, what happened? Did formerly mobile people settle in first in
near-coastal areas, or was there substantial backflow of population into the inte-
rior? Was the wave uniform, or was its shape contorted by variations in topography
and faunal distribution? It is intriguing to view the distribution of known fluted
points (Anderson and Faught, 1998) with the Martin model in mind. Point con-
centrations appear to form a nearly continuous arc, stretching across the southern
United States from California to Maine, as though the outermost concentric rip-
ple of the blitzkrieg wave had frozen in stone. This pattern could be interpreted
as the result of the dense wave-front population beginning to settle in after en-
countering barriers to further radial movement [or, as Faught (1996) proposes, it
might result from drowning of the continental shelf and displacement of its former
human inhabitants]. The point distribution shows no obvious relationship to envi-
ronmental parameters, as the concentrations encompass several distinct vegetation
zones. Thus it does not appear that Paleoindians were searching for any particular
preferred habitat.
If the Clovis drought (Haynes, 1991b) (probably corresponding to the warm
spike at the end of the Allerd, ca. 13,100 B.P.), caused fragmentation of mam-
moth herds into oasis-like refugia, a leap-frogging, long-distance search and de-
stroy strategy, moving from one refugium to the next, might have been most effi-
cient. Thus there may be a processual similarity in the island hopping of the early
Polynesian Lapita people, the movement of Linearbandkeramik people between
patches of arable soil in central Europe, and Clovis migration between dispersed
mammoth refugia, which would explain the seemingly explosive character of each
of these population movements.
Some eastern Paleoindian sites, interpreted as encampments on the basis of
toolkit diversity, are quite large. This site type was, until recently, thought to be re-
stricted to the Northeast (e.g., Bull Brook, Debert, Vail), but large, diverse sites are
now also recognized in the Southeast [e.g., Carson-Conn-Short in Tennessee and
Adams in Kentucky (Anderson, 1996)]. Anderson (1990) has suggested that ini-
tial Paleoindian colonists, moving eastward from the ice-free corridor, would soon
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 83

have encountered the Mississippi. From the central Mississippi valley, they could
have moved east through the valleys of the Ohio, Cumberland, or Tennessee Rivers.
He suggests that these ecologically rich valleys were used as staging areas. Simi-
larly, Dincauze (1993) has seen the large sites as marshalling camps, from which
exploratory forays were mounted into unknown territory. In that case, these large
sites may represent the very first human occupations in eastern North America.
However, it is not obvious exactly how one would distinguish the archaeologi-
cal signature of a marshalling camp set up by the first colonists, from that of a
macroband rendezvous campsite established centuries later. One might predict that
initial camps would be characterized by high percentages of lithic material derived
from the migrants original home territory, but as Spiess et al. (1998) observe in
New England, even at the presumably earliest sites such as Bull Brook, there is ev-
idence of a broad knowledge of regional lithic sources. The region likely had been
thoroughly investigated by scouting parties before the entire social unit moved in
en masse.
The image of squads of flint-knapping specialists reconnoitering new territo-
ries for high-quality stone may be a bit hard to accept, but it is even more difficult
to believe that Paleoindians just stumbled serendipitously upon veins of workable
stone while they were pursuing game or looking for drinking water. It does not
seem that the high-quality lithics were really necessary, from a practical point
of view. Serviceable Clovis points could be made out of quartzite or argillite (as
they were occasionally in southern Virginia); in Tierra del Fuego, Paleoindians
made fishtail points of basalt. Nevertheless, the Clovis culture must have put con-
siderable emphasis on the acquisition of beautiful stones. Fluting, too, was not
necessarily the most practical technique for thinning bifaces. Alternative methods
might have been less costly in terms of accidental breakage. Importantly, these
aspects of Clovis lithic technology may be telling us something about broader cul-
tural values (Storck, 1991). This was perhaps a rather flamboyant culture, investing
a lot of time and energy in nonutilitarian activities. They were flexible enough to
adapt to numerous geographically variable, and rapidly changing, environments,
yet cautiously retained a toolkit of proven efficacy as they traversed thousands of
kilometers.

TOPICS FOR FUTURE RESEARCH

In the coming decades, we can anticipate progress on several fronts in


Paleoindian research. New molecular-scale techniques for analyzing ancient bone
and lithic artifacts have already yielded remarkable results and promise future
surprises. Analysis of skeletal mtDNA (e.g., Stone and Stoneking, 1998) offers a
new way to establish population relationships and perhaps will lead to refined es-
timation of mutation rates. Unfortunately, if the deep ideological divide between
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

84 Fiedel

archaeologists and native people concerning study of North American human


remains cannot be bridged, and indiscriminate repatriation of ancient skeletons
such as Buhl Woman and Kennewick Man continues, we can expect that future
progress in genetic and skeletal research will be based mainly on South American
specimens.
Pioneer immunological studies of purported blood or protein residues on
Paleoindian lithic artifacts yielded intriguing results (e.g., Gramly, 1991; Hyland
et al., 1990; Loy and Dixon, 1998), but where faunal remains are available for
comparison, blood residue findings have sometimes appeared implausible, and
reported results have not been replicated (Eisele et al., 1996; Fiedel, 1996b). Clear
successes with later, more common artifacts must be demonstrated before any
more Paleoindian artifacts are subjected to extractive laboratory procedures that
may leave behind no usable samples for researchers of the future. Phytoliths or
starch grains might be present on Paleoindian tools, particularly scrapers, which
could shed light on use of plants. Recognizable ancient pig DNA is reported to
have been found on a ca. 35,000-year-old Mousterian tool (Hardy et al., 1997);
under the right conditions, DNA might be recovered from Paleoindian tools (as
already claimed by Loy and Dixon, 1998).
Recent correction of radiocarbon dates has shown that Clovis expansion oc-
curred about 13,50013,000 B.P. Plateaus and sharp dips in 14 C between 14,700
and 11,200 B.P. (12,50010,000 rcbp) complicate dating of cultural developments
in this period. For example, it may be impossible, because of a plateau, to discrim-
inate between assemblages from 11,700 and 12,400 B.P.; both would yield radio-
carbon dates of ca. 10,300 rcbp (Bjorck et al., 1996; Burr et al., 1998). Continuing
refinement of thermoluminescence and optical luminescence methods (Feathers,
1997), which may permit accurate dating of burned chert and wind-blown sand
grains, respectively, may provide an alternative route toward fine-tuning Paleoin-
dian chronology.
Also important for this goal are discovery and excavation of more stratified
sites. Such sites are most likely to occur within deep alluvial deposits in river val-
leys. For example, the Big Eddy site, on the Sac River in southwest Missouri, has
yielded Dalton material from a depth of 2.8 to 3.2 m below surface, a Gainey-style
Clovis assemblage (Gainey is an early fluted point type in the Great LakesMidwest
region) at 3.33.5 m, and nondiagnostic lithics below this Clovis level. Published
dates are 10,185 75 for the Dalton horizon, 10,710 85 for the ClovisGainey
horizon, and 10,940 80 for the untyped lithics (Ray et al., 1998); nine unpub-
lished dates for the Clovis level cluster between 11,350 and 10,900 rcbp (Lopinot,
personal communication, 1998). Three stratified Clovis horizons are reportedly
present at the CarsonConnShort site in Tennessee (Broster and Norton, 1996;
Stanford, 1998). As Meadowcroft has shown, Paleoindian material also may be
deeply buried in rockshelters. Late Paleoindian lithics and associated floral and
faunal material are reported from the lowest strata of Dust Cave, Alabama, dated
by several radiocarbon assays to ca. 10,300 rcbp (Driskell, 1996).
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 85

Wide acceptance of the validity of the ca. 12,50012,000 rcbp age of Monte
Verde has sparked a revival of the quest for pre-Clovis sites in North America.
Based on environmental considerations (open vegetation, relatively mild winters,
and high animal biomass), Butzer (1991) has suggested that the areas of North
America most likely to yield evidence of 40,00015,000 rcbp occupation would
be eastern Washington and Oregon, valleys in the desert Southwest and interior
Mountain West, and the middle Appalachians and Mid-Atlantic coast. This last
region may actually have yielded such a siteCactus Hill in southeastern Virginia,
with dates of 15,000 and 16,000 rcbp for an ostensibly pre-Clovis blade assem-
blage. The similarity of the Cactus Hill artifacts to the Meadowcroft assemblage,
with its controversial dates of ca. 14,00014,500 rcbp, lends support to the idea
that people with a Nenana-like Upper Paleolithic toolkit may have been in the
East by 14,000 rcbp (17,000 B.P.). This conclusion is problematic because it re-
quires a migration from Beringia preceding the Blling warming at 12,500 rcbp
(14,700 B.P.).
Given that genetic data and geographic logic continue to point to Northeast
Asia as the Paleoindian ancestral homeland, ongoing and future research in North
China, Mongolia, and Siberia will surely clarify the origin of American cultures.
The chronology and relationships of early cultures in Alaska also will be eluci-
dated. Recent research indicates that the Nenana complex was in place by about
11,800 rcbp (i.e., between 14,000 and 13,500 B.P.). The age of the typically small
fluted points in northern Alaska remains uncertain (Clark, 1991; Reanier, 1995).
Rather than a backwash from the northern Plains, they may yet prove to be the
earliest, ancestral form, perhaps derived from the triangles that typify Nenana
assemblages.
It has long been suspected that the Mid-South contains the highest concen-
tration of fluted points in the continent. This impression is confirmed by recent
work of David Anderson and Michael Faught (1998), who have been collecting
data on the continent-wide distributions of fluted points. A sample of 12,163 points
(including 1465 Folsom and 256 Cumberland points) has been tabulated thus far
(Anderson et al., 1998). The results show a lopsided prevalence of Clovis points in
the East, although there is an unexpected concentration in southcentral California.
However, the typical occurrence of Paleoindian artifacts as isolated surface finds
has discouraged intensive investigation in the South. The recent exciting finds at
Cactus Hill, Big Pine Tree, Dust Cave, and CarsonConnShort suggest that the
South has more to offer, and the region is assuming its rightful central role in
Paleoindian research (Anderson et al., 1996). Continued underwater excavations
in Florida rivers, sinkholes, and offshore deposits (Carter, 1998; Dunbar and Webb,
1996; Faught, 1996; Muniz, 1998) promise either to validate or to dismiss the few
ostensibly pre- or early Clovis materials, dated ca. 12,000 rcbp, from Little Salt
Springs and Page-Ladson.
Very important work is being conducted by Argentine archaeologists, gen-
erally on shoestring budgets, on Fell I fishtail point sites in Tierra del Fuego,
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

86 Fiedel

Patagonia, and the Pampas (Fig. 2) (Flegenheimer, 1986/1987; Nami, 1994, 1997;
Politis, 1997). Dates ranging from about 11,100 to 10,300 rcbp show the basic
contemporaneity of this horizon with Clovis. Further discoveries and analyses of
Central American Paleoindian sites (e.g., Morrow and Morrow, 1997; Pearson,
1998) should clarify and date the evident stylistic transition that occurred there,
from Clovis to fishtail styles. Recovery and dating of more Late Pleistocene faunal
assemblages in South America should show if extinction was as abrupt there as
in the north and if humans were an important factor. Future research in western
and southern South America also should reveal more sites in the range of ca.
13,00011,500 rcbp (16,00013,500 B.P.); if Monte Verde is really as old as it
seems, it should not be unique. More work also is needed in Brazil to explain
the ca. 12,000 rcbp dates reported from a few sites (Kipnis, 1998; Prous and
Fogaca, 1999). Even if the oldest secure dates there prove to be about 11,000 rcbp
(Roosevelt et al., 1996), ostensible contemporaneity with Clovis will still require
an explanation. Why are there no traces of megafauna at any of these early sites?
Is this a matter of cultural choice, seasonality of activities, or were the megafauna
of the region (e.g., the species represented at Lagoa Santa) already extinct? The
possibility remains that there may be a significant offset between Southern and
Northern Hemisphere radiocarbon dates in the Late Pleistocene, which might ex-
plain why the spread of fishtails down to the Southern Cone appears so incredibly
rapid.
Geological and archaeological field research in Canada, within the ice-free
corridor and at its southern end, has not resolved uncertainties about the feasibility
and possible dates of migration by this route. Stratified sites in this area [Charlie
Lake Cave (Fladmark et al., 1988), Vermilion Lakes (Fedje et al., 1995)] contain
basal deposits no earlier than ca. 10,800 rcbp, probably too late to be relevant to
the question of initial peopling. Work in the northern Plains has produced more
evidence of the problematic Goshen complex, which may be as early as Clovis [but
probably is not, based on credible dates of ca. 10,200 rcbp from the Jim Pitts site
(Donohue, 1998)]. If Paleoindians entered by this route, which logic and theory
still suggest, they must have been traveling fast and did not linger. If there are
remains of a big Clovis marshalling camp somewhere near Edmonton, one would
think they should have turned up by now. The same argument applies, however,
to the alleged coastal route. Despite years of contracted and academic survey
work, neither the Northwest nor the California coast has yet produced a single
credible pre-Clovis site. Drowned Late Pleistocene coastal landscapes have been
recognized at a depth of more than 150 m in the waters off the Queen Charlotte
Islands, and Fedje has recently dredged up a single basalt flake, estimated to date
to ca. 10,200 rcbp, from this context (Hall, 1999b). However, given the likely
cost and hit-or-miss nature of such research, future searches might concentrate
on the coast of mainland British Columbia, where near-coastal deposits of similar
antiquity now sit more than 200 m above the present sea level, because of isostatic
postglacial uplift (Josenhans et al., 1997, p. 74).
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 87

Computer-generated simulation models (e.g., Anderson et al., 1997; Gillam,


1998; Gibbs et al., 1998; Whitney-Smith, 1998; Young and Bettinger, 1997)
promise to move the debates over colonization processes and megafaunal extinc-
tion beyond our former reliance on intuitive judgments of plausibility. By starting
with reasonable assumptions concerning reproduction, movement, and kill rates,
these simulations can show whether rapid migration and overkill are replicable,
and they may even point to likely colonization routes that warrant further inves-
tigations. Ultimately, however, virtual simulations, no matter how realistic, will
never be a substitute for actual hard archaeological evidence.

ACKNOWLEDGMENTS

My thanks go to Vance Haynes, David Anderson, Jon Erlandson, and Vance


Holliday for their editorial and bibliographic suggestions.

REFERENCES CITED

Ackerman, R. E. (1992). Earliest stone industries on the North Pacific coast of North America. Arctic
Anthropology 29(2): 1827.
Adovasio, J. M. (1993). The ones that will not go away: A broad view of pre-Clovis populations in the
New World. In Soffer, O., and Praslov, N. D. (eds.), From Kostenki to Clovis, Plenum Press, New
York, pp. 199218.
Adovasio, J. M., Donahue, J., and Stuckenrath, R. (1990). The Meadowcroft Rockshelter radiocarbon
chronology 19751990. American Antiquity 55: 348354.
Agenbroad, L., Morris, D., Rockwell, T., and Roth, L. (1995). A preliminary report of the recovery
of the first nearly complete skeleton of the pygmy (dwarf ) mammoth, Mammuthus exilis, from
North America. Paper presented at INQUA Congress XIV, Berlin, August 310.
Anderson, D. G. (1990). The Paleoindian colonization of eastern North America: A view from the south-
eastern United States. In Tankersley, K. B., and Isaac, B. L. (eds.), Early Paleoindian Economies
of Eastern North America, Research in Economic Anthropology, Supplement No. 5, JAI Press,
Greenwich, CT, pp. 163216.
Anderson, D. G. (1995). Paleoindian interaction networks in the eastern woodlands. In Nassaney, M. S.,
and Sassaman, K. E. (eds.), Native American Interaction: Multiscalar Analyses and Interpretations
in the Eastern Woodlands, University of Tennessee Press, Knoxville, pp. 126.
Anderson, D. G. (1996). Models of Paleoindian and Early Archaic settlement in the lower Southeast.
In Anderson, D. G., and Sassaman, K. E. (eds.), The Paleoindian and Early Archaic Southeast,
University of Alabama Press, Tuscaloosa, pp. 2957.
Anderson, D. G., and Faught, M. K. (1998). The distribution of fluted Paleoindian projectile points:
Update 1998. Archaeology of Eastern North America 26: 163188.
Anderson, D. G., OSteen, L. D., and Sassaman, K. E. (1996). Environmental and chronological
considerations. In Anderson, D. G., and Sassaman, K. E. (eds.), The Paleoindian and Early
Archaic Southeast, University of Alabama Press, Tuscaloosa, pp. 315.
Anderson, D. G., Faught, M. K., and Gillam, J. C. (1997). Paleoindian colonization of the New World:
Implications from an examination of resource structure, physiography, and demography. Paper
presented at 62nd annual meeting of Society for American Archaeology, Nashville.
Anderson, D. G., Faught, M. K., and Gillam, J. C. (1998). Paleoindian site/artifact distributions viewed
from a very large scale: Evidence and implications. Paper presented at 63rd annual meeting of
Society for American Archaeology, Seattle.
Anderson, J. E. (1967). The human skeleton. In Byers, D. S. (ed.), The Prehistory of the Tehuacan
Valley, Vol. 1, University of Texas Press, Austin, pp. 9497.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

88 Fiedel

Anthony, D. W. (1990). Migration in archeology: The baby and the bathwater. American Anthropologist
92: 895914.
Arriaza, B. (1995). Chiles Chinchorro mummies. National Geographic 187(3): 6889.
Aveleyra A. de Anda, L. (1956). The second mammoth and associated artifacts at Santa Isabel Iztapan,
Mexico. American Antiquity 22: 1228.
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. 209230.
Bellwood, P. (1987). The Polynesians: The Prehistory of an Island People, Thames and Hudson,
London.
Beynon, D. E., and Siegel, M. I. (1981). Ancient human remains from central Peru. American Antiquity
46: 167179.
Bianchi, N. O., Bailliet, G., Bravi, C. M., Carnese, R. F., Rothhammer, F., Martinez-Marignac, V. L.,
and Pena, S. D. (1997). Origin of Amerindian Y-chromosomes as inferred by the analysis of six
polymorphic markers. American Journal of Physical Anthropology 102: 7989.
Bianchi, N. O., Catanesi, C. I., Bailliet, G., Martinez-Marignac, V. L., Bravi, C. M., Vidal-Rioja, L. B.,
Herrera, R. J., and Lopez-Camelo, J. S. (1998). Characterization of ancestral and derived Y-
chromosome haplotypes of New World native populations. American Journal of Human Genetics
63: 18621871.
Bird, J. (1938). Before Magellan. Natural History 41(1): 1628.
Birdsell, J. B. (1968). Some predictions for the Pleistocene based on equilibrium systems among
recent hunter-gatherers. In Lee, R. B., and De Vore, I. (eds.), Man the Hunter, Aldine, Chicago,
pp. 229240.
Bjorck, S., Kromer, B., Johnsen, S., Bennike, O., Hammarlund, D., Lemdahl, G., Possnert, G.,
Rasmussen, T. L., Wohlfarth B., Hammer, C. U., and Spurk, M. (1996). Synchronized terrestrial-
atmosphere deglacial records around the North Atlantic. Science 274: 11551160.
Bonatto, S. L., and Salzano, F. M. (1997). Diversity and age of the four major mtDNA haplogroups,
and their implications for the peopling of the New World. American Journal of Human Genetics
61: 14131423.
Bordes, F. (1968). The Old Stone Age, Weidenfeld and Nicolson, London.
Bradley, B. A. (1997). Bifacial thinning in the early Upper Paleolithic of Eastern Europe. Chips 9(2):
89.
Bradley, B. A., Anikovich, M., and Giria, E. (1995). Early Upper Palaeolithic in the Russian Plain:
Streletskayan flaked stone artefacts and technology. Antiquity 69: 989998.
Brerewood, E. (1622). Enquiries Touching the Diversity of Languages, and Religions, Through the
Chief Parts of the World, John Bill, London.
Breternitz, D. A., Swedlund, A. C., and Anderson, D. C. (1971). An early burial from Gordon Creek,
Colorado. American Antiquity 36: 170182.
Brose, D. S. (1994). Archaeological investigations at the Paleo Crossing site, a Paleoindian occupation
in Medina County, Ohio. In Dancey, W. S. (ed.), The First Discovery of America, Archaeological
Evidence of the Early Inhabitants of the Ohio Area, Ohio Archaeological Council, Columbus,
pp. 6176.
Broster, J. B., and Norton, M. R. (1996). Recent Paleoindian research in Tennessee. In Anderson, D. G.,
and Sassaman, K. E. (eds.), The Paleoindian and Early Archaic Southeast, University of Alabama
Press, Tuscaloosa, pp. 288297.
Brown, M. D., Hosseini, S. H., Torroni, A., Bandelt, H. J., Allen, J. C., Schurr, T. G., Scozzari, R.,
Cruciani, F., and Wallace, D. C. (1998). mtDNA haplogroup X: An ancient link between Eu-
rope/ Western Asia and North America? American Journal of Human Genetics 63: 18521861.
Brown, P. (1999). The earliest East Asians: A view from the Late Pleistocene and Neolithic of China
and Japan. In Omoto, K. (ed.), Interdisciplinary Perspectives on the Origins of the Japanese,
International Research Center for Japanese Studies, Kyoto (in press).
Bryan, A. L. (1978). An overview of Paleo-American prehistory from a circum-Pacific perspective. In
Bryan, A. L. (ed.), Early Man in America from a Circum-Pacific Perspective, Occasional Papers,
No. 1, Department of Anthropology, University of Alberta, Edmonton, pp. 306327.
Bryan, A. L. (1983). South America. In Shutler, R., Jr. (ed.), Early Man in the New World, Sage, Beverly
Hills, pp. 137146.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 89

Bryan, A. L. (1991). The fluted-point tradition in the AmericasOne of several adaptations to Late
Pleistocene American environments. In Bonnichsen, R., and Turnmire, K. L. (eds.), Clovis: Origins
and Adaptations, Center for the Study of the First Americans, Corvallis, OR, pp. 1534.
Bryan, A. L., Casamiquela, R. M., Cruxent, J. M., Gruhn, R., and Ochsenius, C. (1978). An El Jobo
mastodon kill at Taima-Taima, Venezuela. Science 200: 12751277.
Burns, J. A. (1990). Paleontological perspectives on the ice-free corridor. In Agenbroad, L., Mead, J.,
and Nelson, L. (eds.), Megafauna and Man: Discovery of Americas Heartland, Mammoth Site
of Hot Springs, SD, pp. 6166.
Burr, G. S., Beck, J. W., Taylor, F. W., Recy, J., Edwards, R. L., Cabioch, G., Corr`ege, T., Donahue,
D. J., and OMalley, J. M. (1998). A high-resolution radiocarbon calibration between 11,700 and
12,400 calendar years BP derived from 230 Th ages of corals from Espiritu Santo Island, Vanuatu.
Radiocarbon 40: 10931106.
Butzer, K. W. (1991). An Old World perspective on potential mid-Wisconsinan settlement of the
Americas. In Dillehay, T. D., and Meltzer, D. J. (eds.), The First Americans: Search and Research,
CRC Press, Boca Raton, FL, pp. 137156.
Carlson, R. L. (1983). The Far West. In Shutler, R., Jr. (ed.), Early Man in the New World, Sage, Beverly
Hills, pp. 7396.
Carlson, R. L. (1991). Clovis from the perspective of the ice-free corridor. In Bonnichsen, R., and
Turnmire, K. L. (eds.), Clovis: Origins and Adaptations, Center for the Study of the First
Americans, Corvallis, OR, pp. 8190.
Carter, B. (1998). The Late Pleistocene component of the Page/Ladson site, North Florida: Chronol-
ogy and regional integration. Paper presented at 63rd annual meeting of Society for American
Archaeology, Seattle.
Carter, G. F. (1980). Earlier than You Think: A Personal View of Man in America, Texas A&M University
Press, College Station.
Chard, C. (1960). Routes to Bering Strait. American Antiquity 28: 151158.
Chatters, J. C. (1997). Kennewick man. Smithsonian Institution Web Page.
Chauchat, C. (1988). Early hunter-gatherers on the Peruvian coast. In Keatinge, R. W. (ed.), Peruvian
Prehistory, Cambridge University Press, Cambridge, pp. 4146.
Clark, D. W. (1991). The northern (Alaska-Yukon) fluted points. In Bonnichsen, R., and Turnmire, K. L.
(eds.), Clovis: Origins and Adaptations, Center for the Study of the First Americans, Corvallis,
OR, pp. 3548.
Clark, D. W., and Clark, A. M. (1983). Paleo-Indians and fluted points: Subarctic alternatives. Plains
Anthropologist 28(102): 283292.
Clausen, C. J., Cohen, A. D., Emiliani, C., Holman, J. A., and Stipp, J. J. (1979). Little Salt Spring,
Florida: A unique underwater site. Science 203: 609614.
Cockrell, W. A., and Murphy, L. (1978). Pleistocene man in Florida. Archaeology of Eastern North
America 6: 113.
Correal Urrego, G., and van der Hammen, T. (1977). Investigaciones arqueologicas en los abrigos
rocosos del Tequendama, Biblioteca Banco Popular, Bogota.
Correal Urrego, G., and van der Hammen, T. (1979). Investigaciones arqueologicas en los abrigos
rocosos de Nemocon y Sueva, Fundacion de Investigaciones Arqueologicas Nacionales, Banco de
la Republica, Bogota.
Cotter, J. L. (1937). The occurrence of flints and extinct animals in pluvial deposits near Clovis, New
Mexico, pt. IV, Report on the excavations at the gravel pit in 1936. Proceedings of the Philadelphia
Academy of Natural Sciences 89: 116.
Cotter, J. L. (1981). The Upper PaleolithicHowever it got here, its here (can the Middle Paleolithic
be far behind?). American Antiquity 46: 926928.
Cushing, J., Wenner, A. M., Noble, E., and Daily, M. (1986). A groundwater hypothesis for the origin
of fire areas on the northern Channel Islands, California. Quaternary Research 26: 207217.
Custer, J. F., and Stewart, R. M. (1990). Environment, analogy, and early Paleoindian economies in
northeastern North America. In Tankersley, K. B., and Isaac, B. L. (eds.), Early Paleoindian
Economies of Eastern North America, Research in Economic Anthropology, Supplement No. 5,
JAI Press, Greenwich, CT, pp. 303322.
de Acosta, J. (1604). The Natural and Moral History of the Indies, English translation by E. Grimston,
1604, ed. C. R. Markham 1880, Hakluyt Society, London.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

90 Fiedel

de Lumley, H., de Lumley, M. A., Beltrao, M., Yokoyama, Y., Labeyrie, J., Danon, J., Delibrias,
G., Falgueres, C., and Bischoff, J. L. (1988). Decouverte doutils tailles associes a` des faunes
du Pleistoc`ene moyen dans la Toca da Esperanca, Etat de Bahia, Bresil. Comptes Rendus
de lAcademie des Sciences, Paris 306: 242247.
Derevianko, A. P. (1989). The Late Pleistocene sites in the Slendia River Basin and their significance
for correlation with Upper Paleolithic assemblages of the Pacific. In Circum-Pacific Prehistory
Conference Program and Abstracts, Seattle, p. 36.
Derevianko, A. P., Shimkin, D. B., and Powers, W. R. (1998). The Paleolithic of Siberia: New Discov-
eries and Interpretations, University of Illinois Press, Urbana.
Dikov, N. N. (1996). The Ushki sites, Kamchatka Peninsula. In West, F. H. (ed.), American Beginnings:
The Prehistory and Palaeoecology of Beringia, University of Chicago Press, Chicago, pp. 244
250.
Dikov, N. N., and Titov, E. E. (1984). Problems of the stratification and periodization of the Ushki
sites. Arctic Anthropology 21(2): 168.
Dillehay, T. D. (1988). Early rainforest archaeology in southwestern South America: Research context,
design and data at Monte Verde. In Purdy, B. A. (ed.), Wet Site Archaeology, Telford Press,
Caldwell, NJ, pp. 177206.
Dillehay, T. D. (1989). Monte Verde, a Late Pleistocene Settlement in Chile, Vol. 1, Palaeoenvironment
and Site Context, Smithsonian Institution Press, Washington, DC.
Dillehay, T. D. (1997). Monte Verde, a Late Pleistocene Settlement in Chile, Vol. 2. The Archaeological
Context and Interpretation, Smithsonian Institution Press, Washington, DC.
Dillehay, T. D., and Meltzer, D. J. (1991). Finale: Processes and prospects. In Dillehay, T. D., and
Meltzer, D. J. (eds.), The First Americans: Search and Research, CRC Press, Boca Raton, FL,
pp. 287294.
Dincauze, D. (1984). An archaeo-logical evaluation of the case for pre-Clovis occupations. In Wendorf,
F., and Close, A. (eds.), Advances in World Archaeology, Vol. 3, Academic Press, New York,
pp. 275323.
Dincauze, D. (1993). Fluted points in the eastern forests. In Soffer, O., and Praslov, N. D. (eds.),
From Kostenki to Clovis: Upper Paleolithic-Paleoindian Adaptations, Plenum Press, New York,
pp. 279292.
Donohue, J. A. (1998). The chronology of the Jim Pitts site and implications for the temporal placement
of the Goshen-Plainview complex in the northwestern Plains. Paper presented at 63rd annual
meeting of Society for American Archaeology, Seattle.
Dorn, R. I. (1996). Uncertainties in the radiocarbon dating of organics associated with rock varnish: A
plea for caution. Physical Geography 17: 585591.
Driskell, B. N. (1996). Stratified Late Pleistocene and Early Holocene deposits at Dust Cave, north-
western Alabama. In Anderson, D. G., and Sassaman, K. E. (eds.), The Paleoindian and Early
Archaic Southeast, University of Alabama Press, Tuscaloosa, pp. 315330.
Dunbar, J. S., and Webb, S. D. (1996). Bone and ivory tools from submerged Paleoindian sites in
Florida. In Anderson, D. G., and Sassaman, K. E. (eds.), The Paleoindian and Early Archaic
Southeast, University of Alabama Press, Tuscaloosa, pp. 331353.
Eisele, J., Fowler, D. D., Haynes, G., and Lewis, R. A. (1996). Survival and detection of blood residues
on stone tools. Antiquity 69: 3646.
Elias, S. A., Short, S. K., Nelson, C. H., and Birks, H. H. (1996). The life and times of the Bering land
bridge. Nature 382: 6063.
Elston, R. G., Cheng, X., Madsen, D. B., Kan, Z., Bettinger, R. L., Jingzen, L., Brantingham, P. J.,
Huiming, W., and Jun, Y. (1997). New dates for the North China Mesolithic. Antiquity 71(274):
985993.
Erlandson, J. M. (1994). Early Hunter-Gatherers of the California Coast, Plenum Press, New York.
Erlandson, J. M. (1998). Paleocoastal occupations of Daisy Cave, San Miguel Island, California. Paper
presented at 63rd Annual Meeting of the Society for American Archaeology, Seattle.
Erlandson, J. M., Kennett, D. L., Ingram, B. L., Guthrie, D. A., Morris, D. P., Tveskov, M. A., West,
G. J., and Walker, P. L. (1996). An archaeological and paleontological chronology for Daisy Cave
(CA-SMI-261), San Miguel Island, California. Radiocarbon 38: 355373.
Fairbanks, R. G., Charles, C. D., and Wright, J. D. (1992). Origin of global meltwater pulses. In Taylor,
R. E., Long, A., and Kra, R. S. (eds.), Radiocarbon After Four Decades, Springer-Verlag, New
York, pp. 473500.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 91

Faught, M. K. (1996). Clovis Origins and Underwater Prehistoric Archaeology in Northwestern


Florida, Ph.D. dissertation, Department of Anthropology, University of Arizona, Tucson.
Feathers, J. K. (1997). The application of luminescence dating in American archaeology. Journal of
Archaeological Method and Theory 4(1): 166.
Fedje, D. W., White, J. M., Wilson, M. C., Nelson, D. E., Vogel, J. S., and Southon, J. R. (1995).
Vermilion Lakes site: Adaptations and environments in the Canadian Rockies during the latest
Pleistocene and early Holocene. American Antiquity 60: 81108.
Fedje, D. W., McSporran, J. B., and Mason, A. R. (1996). Early Holocene archaeology and paleoecology
at the Arrow Creek sites in Gwaii Haanas. Arctic Anthropology 33(1): 116142.
Ferring, C. R. (1995). The Late Quaternary geology and archaeology of the Aubrey Clovis site, Texas:
A preliminary report. In Johnson, E. (ed.), Ancient Peoples and Landscapes, Museum of Texas
Tech University, Lubbock, pp. 273281.
Fiedel, S. J. (1987). Prehistory of the Americas, Cambridge University Press, Cambridge.
Fiedel, S. J. (1996a). Paleoindians in the Brazilian Amazon. Science 274: 18211822.
Fiedel, S. J. (1996b). Blood from stones? Some methodological and interpretive problems in blood
residue analysis. Journal of Archaeological Science 23: 139147.
Fiedel, S. J. (1998). Rapid migrations by Arctic hunting peoples: Clovis and Thule. Paper presented at
63rd annual meeting of Society for American Archaeology, Seattle.
Fiedel, S. J. (1999). Older than we thought: Implications of corrected dates for Paleoindians. American
Antiquity 64: 95116.
Fiedel, S. J., and Anthony, D. (1979). The diffusion of the early European Neolithic: A reconsideration.
Paper delivered at the Second Eastern European Archaeology (Hleb i Vino) Conference, University
of Pennsylvania, Philadelphia.
Fifield, T. E. (1996). Human remains in Alaska reported to be 9,730 years old. SAA Bulletin 14(5): 5.
Fladmark, K. R. (1979). Routes: Alternate migration corridors for early man in North America.
American Antiquity 44: 5569.
Fladmark, K. R. (1983). Times and places: Environmental correlates of mid-to-late Wisconsinan human
population expansion in North America. In Shutler, R., Jr. (ed.), Early Man in the New World,
Sage, Beverly Hills, pp. 1342.
Fladmark, K. R., Driver, J. C., and Alexander, D. (1988). The Paleoindian component at Charlie Lake
Cave (HbRf 39), British Columbia. American Antiquity 53: 371384.
Flannery, T. F. (1999). Debating extinction. Science 283: 182183.
Flegenheimer, N. (1986/1987). Excavaciones en el sitio 3 de la localidad Cerro La China (Prov. de Bs.
As.). Relaciones de la Sociedad Argentina de Antropologa XVII-I: 728.
Flegenheimer, N., and Zarate, M. (1997). Considerations on radiocarbon and calibrated dates from
Cerro La China and Cerro El Sombrero, Argentina. Current Research in the Pleistocene 14:
2728.
Forster, P., Harding, R., Torroni, A., and Bandelt, H.-J. (1996). Origin and evolution of Native
American mtDNA variation: A reappraisal. American Journal of Human Genetics 59: 935
945.
Frison, G. C. (1987). The Goshen Paleoindian complex: New data for Clovis and Folsom research.
Paper presented at the Twelfth International Congress of the International Union for Quaternary
Research, Ottawa, Canada.
Frison, G. C. (1988). Paleoindian subsistence and settlement during post-Clovis times on the northwest-
ern Plains, the adjacent mountain ranges, and intermontane basins. In Carlisle, R. C. (ed.), America
Before Columbus: Ice Age Origins, Ethnology Monographs No. 12, University of Pittsburgh,
Pittsburgh, pp. 83106.
Frison, G. C. (1991). The Goshen Paleoindian complex: New data for Paleoindian research. In
Bonnichsen, R., and Turnmire, K. L. (eds.), Clovis: Origins and Adaptations, Center for the
Study of the First Americans, Corvallis, OR, pp. 133152.
Frison, G. C. (1993). The North American Paleoindian: A wealth of new data but still much to learn.
Plains Anthropologist 38(145), Memoir 27, pp. 516.
Frison, G. C. (1996). The Mill Iron Site, University of New Mexico Press, Albuquerque.
Gardner, W. M. (1974). The Flint Run complex: Pattern and process during the Paleoindian to Early
Archaic. In The Flint Run Complex: A Preliminary Report 197173 Seasons, Occasional Publi-
cation, No. 1, Archeology Laboratory, Department of Anthropology, The Catholic University of
America, Washington, DC, pp. 547.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

92 Fiedel

Gardner, W. M. (1989). An examination of cultural change in the Late Pleistocene and Early Holocene
(circa 9200 to 6800 B.C.). In Wittkofski, J. M., and Reinhart, T. R. (eds.), Paleoindian Research in
Virginia: A Synthesis, Special Publication, No. 19, Archeological Society of Virginia, Richmond,
pp. 551.
Gibbs, K., Glass, C., and Steele, J. (1998). Calculating best estimates of rates, routes, and times of
earliest Paleoindian dispersals into South America. Paper presented at 63rd Annual Meeting of
the Society for American Archaeology, Seattle.
Gillam, J. C. (1998). Hemispheric-scale modeling of Paleoindian migration. Paper presented at 63rd
Annual Meeting of the Society for American Archaeology, Seattle.
Goddard, I., and Campbell, L. (1994). The history and classification of American Indian languages:
What are the implications for the peopling of the Americas? In Bonnichsen, R., and Steele, D. G.
(eds.), Method and Theory for Investigating the Peopling of the Americas, Center for the Study of
the First Americans, Corvallis, OR, pp. 189208.
Goebel, T. (1993). Characterizing the Siberian Middle-Upper Paleolithic transition. Paper presented at
58th Annual Meeting of the Society for American Archaeology, St. Louis.
Goebel, T. (1998). The Siberian Upper Paleolithic: Hard environments, limiting factors, and human
range expansion into northern Asia. Paper presented at 63rd Annual Meeting of the Society for
American Archaeology, Seattle.
Goebel, T., and Aksenov, M. (1995). Accelerator radiocarbon dating of the initial Upper Palaeolithic
in southeastern Siberia. Antiquity 69: 349357.
Goebel, T., Powers, R., and Bigelow, N. (1991). The Nenana complex of Alaska and Clovis origins.
In Bonnichsen, R., and Turnmire, K. L. (eds.), Clovis: Origins and Adaptations, Center for the
Study of the First Americans, Corvallis, OR, pp. 4980.
Goodyear, A. C. (1997). Clovis utilization at southeastern U.S. quarries: The Big Pine Tree site,
Allendale, South Carolina. Paper presented at 62nd Annual Meeting of the Society for American
Archaeology, Nashville.
Goodyear, A. C., Foss, J., Wah, J., and Wagner, G. (1998). Evidence of pre-Clovis lithic remains
in Allendale County, South Carolina. Paper presented at annual meeting of the Southeastern
Archeological Conference, Greenville, SC.
Graham, R. (1998). Mammals eye view of environmental change in the United States at the end of the
Pleistocene. Paper presented at 63rd Annual Meeting of the Society for American Archaeology,
Seattle.
Gramly, R. M. (1991). Blood residues upon tools from the East Wenatchee Clovis site, Douglas County,
Washington. Ohio Archaeologist 41(4): 49.
Gramly, R. M., and Funk, R. E. (1990). What is known and not known about the human occupation
of the northeastern United States until 10,000 B.P. Archaeology of Eastern North America 18: 5
32.
Grayson, D. K. (1991). Late Pleistocene mammalian extinction in North America: Taxonomy, chronol-
ogy, and explanations. Journal of World Prehistory 5: 193231.
Green, F. E. (1963). The Clovis blades: An important addition to the Llano complex. American Antiquity
29: 145165.
Green, T. J., Cochran, B., Fenton, T. W., Woods, J. C., Titmus, G. L., Tieszen, L., Davis, M. A.,
and Miller, S. J. (1998). The Buhl burial: A Paleoindian woman from southern Idaho. American
Antiquity 63: 437456.
Greenberg, J. H. (1987). Language in the Americas, Stanford University Press, Palo Alto, CA.
Gruhn, R. (1988). Linguistic evidence in support of the coastal route of earliest entry into the New
World. Man 23: 77100.
Gruhn, R. (1991). Stratified radiocarbon-dated archaeological sites of Clovis age and older in Brazil.
In Bonnichsen, R., and Turnmire, K. L. (eds.), Clovis: Origins and Adaptations, Center for the
Study of the First Americans, Corvallis, OR, pp. 283286.
Gruhn, R. (1994). The Pacific coast route of initial entry: An overview. In Bonnichsen, R., and Steele,
D. G. (eds.), Method and Theory for Investigating the Peopling of the Americas, Center for the
Study of the First Americans, Corvallis, OR, pp. 249256.
Guidon, N., and Delibrias, G. (1986). Carbon-14 dates point to man in the Americas 32,000 years ago.
Nature 321: 769771.
Gustafson, C., Daugherty, R., and Gilbow, D. (1979). The Manis Mastodon site: Early man on the
Olympic Peninsula. Canadian Journal of Archaeology 3: 157164.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 93

Hall, D. A. (1997). 8,000-year-old cave explorer challenges research team. Mammoth Trumpet 12(1):
1011.
Hall, D. A. (1999a). Explaining Pleistocene extinctions. Mammoth Trumpet 14(1): 1423.
Hall, D. A. (1999b). Charting the way into the Americas. Mammoth Trumpet 14(1): 111.
Hardy, B. L., Raff, R. A., and Raman, V. (1997). Recovery of mammalian DNA from Middle Paleolithic
stone tools. Journal of Archaeological Science 25: 177184.
Haury, E. W., Sayles, E. B., and Wasley, W. W. (1959). The Lehner Mammoth site, southeastern
Arizona. American Antiquity 25: 230.
Hauswirth, W. W., Dickel, C. D., Rowold, D. J., and Hauswirth, M. A. (1994). Inter- and intrapopulation
studies of ancient humans. Experientia 50: 585591.
Hayden, B. (1982). Interaction parameters and the demise of Paleo-Indian craftsmanship. Plains An-
thropologist 27: 109123.
Haynes, C. V., Jr. (1966). Elephant-hunting in North America. Scientific American 214(6): 104112.
Haynes, C. V., Jr. (1974). Paleoenvironments and cultural diversity in late Pleistocene South America:
A reply to A. L. Bryan. Quaternary Research 4: 378382.
Haynes, C. V., Jr. (1980a). The Clovis culture. Canadian Journal of Anthropology 1(1): 115122.
Haynes, C. V., Jr. (1980b). Paleoindian charcoal from Meadowcroft Rock Shelter: Is contamination a
problem? American Antiquity 45: 583587.
Haynes, C. V., Jr. (1982). Were Clovis progenitors in Beringia? In Hopkins, D. M., Matthews, J. V.,
Schweger, C. E., and Young, S. B. (eds.), Paleoecology of Beringia, Academic Press, New York,
pp. 383398.
Haynes, C. V., Jr. (1987). Clovis origin update. The Kiva 52: 8393.
Haynes, C. V., Jr. (1991a). More on Meadowcroft radiocarbon chronology. Review of Archaeology
12(1): 814.
Haynes, C. V., Jr. (1991b). Geoarchaeological and paleohydrological evidence for a Clovis-age drought
in North America and its bearing on extinction. Quaternary Research 35: 438450.
Haynes, C. V., Jr. (1992). Contributions of radiocarbon dating to the geochronology of the peopling of
the New World. In Taylor, R. E., Long, A., and Kra, R. S. (eds.), Radiocarbon After Four Decades,
Springer-Verlag, New York, pp. 355374.
Haynes, C. V., Jr. (1993). Clovis-Folsom geochronology and climatic change. In Soffer, O., and Praslov,
N. D. (eds.), From Kostenki to Clovis: Upper Paleolithic-Paleoindian Adaptations, Plenum Press,
New York, pp. 219236.
Haynes, C. V., Jr. (1998). Geochronology of the stratigraphic manifestation of paleoclimatic events at
Paleoindian sites. Paper presented at 63rd Annual Meeting of the Society for American Archae-
ology, Seattle.
Haynes, C. V., Jr., Donahue, D. J., Jull, A. J. T., and Zabel, T. H. (1984). Application of accelerator
dating to fluted point Paleoindian sites. Archaeology of Eastern North America 12: 184191.
Helm, J. (1968). The nature of Dogrib socioterritorial groups. In Lee, R. B., and DeVore, I. (eds.), Man
the Hunter, Aldine, Chicago, pp. 118125.
Hoffecker, J. F., Powers, W. R., and Goebel, T. (1993). The colonization of Beringia and the peopling
of the New World. Science 259: 4653.
Holmes, C. E. (1996). Broken Mammoth. In West, F. H. (ed.), American Beginnings: The Palaeoecology
and Prehistory of Beringia, University of Chicago Press, Chicago, pp. 312317.
Holmes, C. E., vanderHoek, R., and Dilley, T. E. (1996). Swan Point. In West, F. H. (ed.), American
Beginnings: The Palaeoecology and Prehistory of Beringia, University of Chicago Press, Chicago,
pp. 319323.
Howell, N., Kubacka, I., and Mackey, D. A. (1996). How rapidly does the human mitochondrial genome
evolve? American Journal of Human Genetics 59: 501509.
Howells, W. (1967). Mankind in the Making, Doubleday, Garden City, NY.
Hyland, D. C., Tersak, J. M., Adovasio, J. M., and Siegel, M. I. (1990). Identification of species of
origin of residual blood on lithic material. American Antiquity 55: 104111.
Ikawa-Smith, F. (1978). Lithic assemblages from the early and middle Upper Paleolithic formations in
Japan. In Bryan, A. L. (ed.), Early Man in America from a Circum-Pacific Perspective, Occasional
Papers, No. 1, Department of Anthropology, University of Alberta, Edmonton, pp. 4253.
Irwin-Williams, C. (1967). Association of early man with horse, camel and mastodon at Hueyatlaco,
Valsequillo (Puebla, Mexico). In Martin, P. S. (ed.), Pleistocene Extinctions, Yale University Press,
New Haven, CT, pp. 337347.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

94 Fiedel

Irwin-Williams, C. (1978). Summary of archaeological evidence from the Valsequillo region, Puebla,
Mexico. In Browman, D. L. (ed.), Cultural Continuity in Mesoamerica, Mouton, London, pp. 7
22.
Irwin-Williams, C. (1981). Letter: Commentary on geologic evidence for age of deposits at Hueyatlaco
archaeological site, Valsequillo, Mexico. Quaternary Research 16: 258.
Jackson, L. E., Jr., and Duk-Rodkin, A. (1996). Quaternary geology of the ice-free corridor: Glacial
controls on the peopling of the New World. In Akazawa, T., and Szathmary, E. J. E. (eds.),
Prehistoric Mongoloid Dispersals, Oxford University Press, Oxford, pp. 214227.
Jantz, R. L., and Owsley, D. W. (1997). Pathology, taphonomy, and cranial morphometrics of the Spirit
Cave mummy. Nevada Historical Society Quarterly 40(1): 6284.
Jennings, J. D. (1974). Prehistory of North America, 2nd ed., McGraw-Hill, New York.
Jennings, J. D. (1983). Ancient North Americans, Freeman, San Francisco.
Johnson, D. L. (1983). The California continental borderland: Landbridges, watergaps and biotic
dispersals. In Masters, P. M., and Fleming, N. C. (eds.), Quaternary Coastlines and Marine
Archaeology: Towards the Prehistory of Land Bridges and Continental Shelves, Academic Press,
London, pp. 481528.
Johnson, M. F. (1997). Confirmation of McAvoys early Cactus Hill sequence. Paper presented at
Middle Atlantic Archaeological Conference, Ocean City, MD.
Jones, T. L. (1991). Marine-resource value and the priority of coastal settlement: A California perspec-
tive. American Antiquity 56: 419443.
Josenhans, H., Fedje, D., Pienitz, R., and Southon, J. (1997). Early humans and rapidly changing
Holocene sea levels in the Queen Charlotte Islands-Hecate Strait, British Columbia, Canada.
Science 277: 7174.
Kamminga, J., and Wright, R. V. S. (1988). The Upper Cave at Zhoukoudian and the origins of the
Mongoloids. Journal of Human Evolution 17: 739767.
Karafet, T. M., Zegura, S. L., Posukh, O., Osipova, L., Bergen. A., Long, J., Goldman, D., Klitz, W.,
Harihara, S., de Knijff, P., Griffiths, R. C., Templeton, A. R., and Hammer, M. F. (1999). Ancestral
Asian source(s) of New World Y-chromosome founder haplotypes. American Journal of Human
Genetics 64: 817831.
Keefer, D. K., deFrance, S. D., Moseley, M. E., Richardson, J. B., III, Satterlee, D. R., and Day-Lewis,
A. (1998). Early maritime economy and El Nino events at Quebrada Tacahuay, Peru. Science
281(5384): 18331835.
Kelly, R. L., and Todd, L. C. (1988). Coming into the country: Early Paleoindian hunting and mobility.
American Antiquity 53: 231244.
King, M. B., and Slobodin, S. B. (1996). A fluted point from the Uptar site, northeastern Siberia.
Science 273: 634636.
Kipnis, R. (1998). Early hunter-gatherers in the Americas: Perspectives from central Brazil. Antiquity
72: 581592.
Kiryak, M. A. (1996). Bolshoi Elgakhchan 1 and 2, Omolon River basin, Magadan district. In West,
F. H. (ed.), American Beginnings: The Prehistory and Palaeoecology of Beringia, University of
Chicago Press, Chicago, pp. 228236.
Kolman, C. J., Sambuughin, N., and Bermingham, E. (1996). Mitochondrial DNA analysis of
Mongolian populations and implications for the origin of New World founders. Genetics 142:
13211334.
Krieger, A. D. (1964). Early man in the New World. In Jennings, J. D., and Norbeck, E. (eds.),
Prehistoric Man in the New World, University of Chicago Press, Chicago, pp. 2384.
Kunz, M. L., and Reanier, R. E. (1994). Paleoindians in Beringia: Evidence from Arctic Alaska. Science
263: 660662.
Kuzmin, Y. V., and Orlova, L. A. (1998). Radiocarbon chronology of the Siberian Paleolithic. Journal
of World Prehistory 12: 153.
Kuzmin, Y. V., and Tankersley, K. B. (1996). The colonization of eastern Siberia: An evaluation of the
Paleolithic age radiocarbon dates. Journal of Archaeological Science 23: 577585.
Kuzmin, Y. V., Jull, A. J. T., Lapshina, Z. S., and Medvedev, V. E. (1997). Radiocarbon AMS dating of
the ancient sites with earliest pottery from the Russian Far East. Nuclear Instruments and Methods
in Physics Research B 123: 496497.
Lahr, M. M. (1995). Patterns of modern human diversification: Implications for Amerindian origins.
Yearbook of Physical Anthropology 38: 163198.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 95

Lewis, T. M. N., and Lewis, M. K. (1961). Eva: An Archaic Site, University of Tennessee Press,
Knoxville.
Lorenz, J. G., and Smith, D. G. (1996). Distribution of four founding mtDNA haplogroups among
native North Americans. American Journal of Physical Anthropology 101: 307323.
Loy, T. H., and Dixon, E. J. (1998). Blood residues on fluted points from eastern Beringia. American
Antiquity 63: 2146.
Lozhkin, A. V., Anderson, P. M., Eisner, W. R., Ravako, L. G., Hopkins, D. M., Brubaker, L. B.,
Colinvaux, P. A., and Miller, M. C. (1993). Late Quaternary pollen records from southwestern
Beringia. Quaternary Research 39: 314324.
Lynch, T. F. (1990). Glacial-age man in South America? A critical review. American Antiquity 55:
1236.
MacNeish, R. S. (1983). Mesoamerica. In Shutler, R., Jr. (ed.), Early Man in the New World, Sage,
Beverly Hills, pp. 125136.
Mandryk, C. A. S. (1996). Late-Glacial vegetation and environment on the eastern slope foothills of
the Rocky Mountains, Alberta, Canada. Journal of Paleolimnology 16: 3757.
Mandryk, C. A. S. (1998). Evaluating paleoenvironmental constraints on interior and coastal entry
routes into North America. Paper presented at 63rd annual meeting of Society for American
Archaeology.
Martin, P. S. (1973). The discovery of America. Science 179: 969974.
Mason, R. J., and Irwin, C. (1960). An Eden-Scottsbluff burial in northeastern Wisconsin. American
Antiquity 26: 4357.
Mayer-Oakes, W. J. (1984). Fluted projectile points-A North American shibboleth viewed in South
American perspective. Archaeology of Eastern North America 12: 231247.
McAvoy, J. M. (1992). Nottoway River Survey, Part I, Clovis Settlement Patterns, Special Publication,
No. 28, Archeological Society of Virginia.
McAvoy, J. M. (1997). A culture sequence from the stratified Cactus Hill site in Sussex County, Virginia.
Paper presented at Middle Atlantic Archaeological Conference, Ocean City, MD.
McGhee, R. (1984). Thule prehistory of Canada. In Damas, D. (ed.), Arctic, Handbook of North
American Indians, Vol. 5, Smithsonian Institution Press, Washington, DC, pp. 369376.
McNett, C. W. (ed.) (1985). Shawnee Minisink: A Stratified Paleoindian-Archaic Site in the Upper
Delaware Valley of Pennsylvania, Academic Press, New York.
Mead, J. I. (1980). Is it really that old? A comment about the Meadowcroft Rockshelter overview.
American Antiquity 45: 579581.
Meighan, C. W. (1983). Early man in the New World. In Masters, P. M., and Fleming, N. C. (eds.),
Quaternary Coastlines and Marine Archaeology: Towards the Prehistory of Land Bridges and
Continental Shelves, Academic Press, London, pp. 441462.
Meltzer, D. J. (1991). On paradigms and paradigm bias in controversies over human antiquity in
America. In Dillehay, T. D., and Meltzer, D. J. (eds.), The First Americans: Search and Research,
CRC Press, Boca Raton, FL, pp. 1352.
Meltzer, D. J. (1994). The discovery of deep time: A history of views on the peopling of the
Americas. In Bonnichsen, R., and Steele, D. G. (eds.), Method and Theory for Investigating
the Peopling of the Americas, Center for the Study of the First Americans, Corvallis, OR, pp. 7
26.
Meltzer, D. J. (1995). Clocking the first Americans. Annual Review of Anthropology 24: 2145.
Meltzer, D. J. (1997). Monte Verde and the Pleistocene peopling of the Americas. Science 276: 754
755.
Meltzer, D. J., Adovasio, J. M., and Dillehay, T. D. (1994). On a Pleistocene human occupation at
Pedra Furada, Brazil. Antiquity 68(261): 695714.
Merriwether, D. A., Rothhammer, F., and Ferrell, R. E. (1994). Genetic variation in the New World:
Ancient teeth, bone, and tissue as sources of DNA. Experientia 50: 592601.
Mithen, S. (1996). The Prehistory of the Mind, Thames and Hudson, London.
Mochanov, Y. A., and Fedoseeva, S. A. (1996a). Dyuktai Cave. In West, F. H. (ed.), American Be-
ginnings: The Prehistory and Palaeoecology of Beringia, University of Chicago Press, Chicago,
pp. 164171.
Mochanov, Y. A., and Fedoseeva, S. A. (1996b). Berelekh, Allakhovsk region. In West, F. H. (ed.),
American Beginnings: The Prehistory and Palaeoecology of Beringia, University of Chicago
Press, Chicago, pp. 218221.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

96 Fiedel

Mochanov, Y. A., and Fedoseeva, S. A. (1996c). Khukhtuy 3. In West, F. H. (ed.), American Beginnings:
The Prehistory and Palaeoecology of Beringia, University of Chicago Press, Chicago, pp. 224
227.
Morell, V. (1998). Kennewick Mans contemporaries. Science 280(5361): 191.
Morlan, R. E. (1991). Peopling of the New World: A discussion. In Bonnichsen, R., and Turnmire, K. L.
(eds.), Clovis: Origins and Adaptations, Center for the Study of the First Americans, Corvallis,
OR, pp. 303308.
Morlan, R. E., and Cinq-Mars, J. (1982). Ancient Beringians: Human occupation in the Late Pleistocene
of Alaska and the Yukon Territory. In Hopkins, D. M. Matthews, J. V., Jr., Schweger, C. E., and
Young, S. B. (eds.), Paleoecology of Beringia, Academic Press, New York, pp. 353381.
Morrow, J., and Morrow, T. (1997). Geographic patterning in fluted point morphology in the New World.
Paper presented at 62nd Annual Meeting of the Society for American Archaeology, Nashville.
Muller-Beck, H. (1966). Paleohunters in America: Origins and diffusion. Science 152: 11911210.
Muniz, M. (1998). Investigations at Little River Rapids (8Je603): A prehistoric inundated site chroni-
cling the Pleistocene/ Holocene transition in north Florida. Paper presented at 63rd Annual Meeting
of the Society for American Archaeology, Seattle.
Nami, H. G. (1994). Resena sobre los avances de la arqueologa finipleistocenica del extremo sur de
Sudamerica. Revista Chungara 26: 145163.
Nami, H. G. (1996). New assessments of early human occupations in the Southern Cone. In Akazawa,
T., and Szathmary, E. J. E. (eds.), Prehistoric Mongoloid Dispersals, Oxford University Press,
Oxford, pp. 256269.
Nami, H. G. (1997). Archaeology and landscapes during the Pleistocene-Holocene transition in the
southern tip of South America. Paper presented at 62nd Annual Meeting of the Society for
American Archaeology, Nashville.
Nelson, D. E., Morlan, R. E., Vogel, J. S., Southon, J. R., and Harington, C. R. (1986). New radiocarbon
dates on artifacts from the northern Yukon Territory: Holocene not Upper Pleistocene in age.
Science 232: 749751.
Neumann, G. K. (1952). Archeology and race in the American Indian. In Griffin, J. B. (ed.), Archeology
of Eastern United States, University of Chicago Press, Chicago, pp. 1334.
New York Times (1997). Disease is new suspect in ancient extinctions, April 29.
New York Times (1998). Chilean field yields new clues to peopling of Americas, August 25.
Oakley, K. P., Campbell, B. G., and Molleson, T. I. (1975). Catalogue of Fossil Hominids, Part III.
Americas, Asia, Australia, Trustees of the British Museum, London.
Orr, P. C. (1968). Prehistory of Santa Rosa Island, Santa Barbara Museum of Natural History, Santa
Barbara, CA.
Overstreet, D. F., and Stafford, T. W., Jr. (1997). Additions to a revised chronology for cultural and
non-cultural mammoth and mastodon fossils in the southwestern Lake Michigan basin. Current
Research in the Pleistocene 14: 7071.
Pearson, G. A. (1997a). Paleoindians in the Alaskan interior: Results of the 1996 Moose Creek expe-
dition. Paper presented at 62nd annual meeting of Society for American Archaeology, Nashville.
Pearson, G. A. (1997b). Non-Mongoloid Pleistocene expansions: Old and new ideas on the origins of the
first Americans. Paper presented at 24th annual meeting of Alaska Anthropological Association,
White Horse, Yukon.
Pearson, G. A. (1998). Pan-American Paleoindian dispersals as seen through the lithic reduction strate-
gies and tool manufacturing techniques at the Guardiria site, Turrialba valley, Costa Rica. Paper
presented at 63rd annual meeting of Society for American Archaeology, Seattle.
Pei, G. (1985). Microlithic industries in China. In Rukang, W., and Olsen, J. W. (eds.), Palaeoanthro-
pology and Palaeolithic Archaeology in the Peoples Republic of China, Academic Press, New
York, pp. 225242.
Petersen, K. L., Mehringer, P. J., Jr., and Gustafson C. E. (1983). Late-glacial vegetation and climate
at the Manis Mastodon site, Olympic Peninsula, Washington. Quaternary Research 20: 215
231.
Politis, G. (1997). The Peopling of the Americas viewed from the Southern Cone. Paper presented at
62nd Annual Meeting of the Society for American Archaeology, Nashville.
Pope, G. G. (1984). The antiquity and paleoenvironment of the Asian Hominidae. In Whyte, R. O. I.
(ed.), The Evolution of the East Asian Environment, Center of Asian Studies, University of Hong
Kong, Hong Kong, pp. 922947.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 97

Price, T. D. (1991). The view from Europe: Concepts and questions about Terminal Pleistocene
societies. In Dillehay T. D., and Meltzer, D. J. (eds.), The First Americans: Search and Research,
CRC Press, Boca Raton, FL, pp. 185208.
Prous, A. (1992). Arqueologia Brasileira, Editora UnB., Brasilia.
Prous, A., and Fogaca, E. (1999). Archaeology of the Pleistocene-Holocene boundary in Brazil. Qua-
ternary International 53/54: 2141.
Purdy, B. A. (1991). The Art and Archaeology of Floridas Wetlands, CRC Press, Boca Raton, FL.
Quade, J., Forester, R. M., Pratt, W. L., and Carter, C. (1998). Black mats, spring-fed streams, and Late
Glacial age recharge in the southern Great Basin. Quaternary Research 49: 129148.
Ranere, A. J. (1997). Paleoindian expansion into tropical America: The view from Central America.
Paper presented at 62nd Annual Meeting of the Society for American Archaeology, Nashville.
Ranere, A. J., and Cooke, R. G. (1991). Paleoindian occupation in the Central American tropics. In
Bonnichsen, R., and Turnmire, K. L. (eds.), Clovis: Origins and Adaptations, Center for the Study
of the First Americans, Corvallis, OR, pp. 237254.
Ray, J. H., Lopinot, N. H., Hajic, E. R., and Mandel, R. D. (1998). The Big Eddy site: A multicomponent
Paleoindian site on the Ozark border, southwest Missouri. Plains Anthropologist 43(163): 7381.
Reanier, R. E. (1995). The antiquity of Paleoindian materials in northern Alaska. Arctic Anthropology
32: 3150.
Rensberger, B. (1997). Putting a new face on prehistory: Skeletons suggest Caucasoid early Americans.
Washington Post, April 15.
Rogers, E. S., and Leacock, E. (1981). Montagnais-Naskapi. In Helm, J. (ed.), Handbook of North
American Indians, Vol. 6. Subarctic, Smithsonian Institution, Washington, DC, pp. 169189.
Roosevelt, A. C., Lima da Costa, M., Lopes Machado, C., Michab, M., Mercier, N., Vallada, H.,
Feathers, J., Barnet, W., Imazio da Silveira, M., Henderson, A., Sliva, J., Chernoff, B., Reese,
D. S., Holman, J. A., Toth, N., and Schick, K. (1996). Paleoindian cave dwellers in the Amazon:
The peopling of the Americas. Science 272: 373384.
Sanders, T. N. (1990). Adams: The Manufacturing of Flaked Stone Tools at a Paleoindian Site in
Western Kentucky, Persimmon Press, Buffalo, NY.
Sandweiss, D. H., McInnis, H., Burger, R. L., Cano, A., Ojeda, B., Paredes, R., del Carmen Sandweiss,
M., and Glascock, M. R. (1998). Quebrada Jaguay: Early South American maritime adaptations.
Science 281(5384): 18301832.
Santos, F. R., Pandya, A., Tyler-Smith, C., Pena, S. D. J., 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: 619628.
Schmitz, P. I. (1987). Prehistoric hunters and gatherers of Brazil. Journal of World Prehistory 1:
53126.
Schurr, T. G., Sukernik, R. I., Starikovskaya, Y. B., and Wallace, D. C. (1999). Mitochondrial DNA
variation in Koryaks and Itelmen: Population replacement in the Okhotsk Sea-Bering Sea region
during the Neolithic. American Journal of Physical Anthropology 108: 139.
Seeman, M. F. (1994). Intercluster lithic patterning at Nobles Pond: A case for disembedded pro-
curement among early Paleoindian societies. American Antiquity 59: 273287.
Seielstad, M. T., Minch, E., and Cavalli-Sforza, L. L. (1998). Genetic evidence for a higher female
migration rate in humans. Nature Genetics 20: 278280.
Sellards, E. H. (1952). Early Man in America, University of Texas Press, Austin.
Sheppard, J. C., Wigand, P. E., Gustafson, C. E., and Rubin, M. (1987). A reevaluation of the Marmes
Rockshelter radiocarbon chronology. American Antiquity 52: 118124.
Shields, G. F., Schmiechen, A. M., Frazier, B. L., Redd, A., Voevoda, M. I., Redd, J. K., and Ward, R. H.
(1993). MtDNA sequences suggest a recent evolutionary divergence for Beringian and northern
North American Populations. American Journal of Human Genetics 53: 549562.
Simons, D. B., Shott, M. J., and Wright, H. T. (1984). The Gainey site: Variability in a Great Lakes
Paleo-Indian assemblage. Archaeology of Eastern North America 12: 266279.
Snarskis, M. J. (1979). Turrialba: A Paleoindian quarry and workshop site in eastern Costa Rica.
American Antiquity 44: 110124.
Sowers, T., and Bender, M. (1995). Climate records covering the last deglaciation. Science 269: 210
214.
Spiess, A. E., Wilson, D., and Bradley, J. W. (1998). Paleoindian occupation in the New England-
Maritimes region: Beyond cultural ecology. Archaeology of Eastern North America 26: 201264.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

98 Fiedel

Stafford, T. W., Jr. (1994). Accelerator C-14 dating of human fossil skeletons: Assessing accuracy
and results on New World specimens. In Bonnichsen, R., and Steele, D. G. (eds.), Method and
Theory for Investigating the Peopling of the Americas, Center for the Study of the First Americans,
Corvallis, OR, pp. 4556.
Stanford, D. (1983). Pre-Clovis occupation south of the ice sheets. In Shutler, R., Jr. (ed.), Early Man
in the New World, Sage, Beverly Hills, pp. 6572.
Stanford, D. (1991). Clovis origins and adaptations: An introductory perspective. In Bonnichsen, R.,
and Turnmire, K. L. (eds.), Clovis: Origins and Adaptations, Center for the Study of the First
Americans, Corvallis, OR, pp. 114.
Stanford, D. (1998). The first Americans: A new perspective. Paper presented at 33rd Annual Sympo-
sium of Archeological Society of Maryland, Crownsville.
Steele, D. G., and Powell, J. F. (1992). Peopling of the Americas: Paleobiological evidence. Human
Biology 64: 303336.
Steele, D. G., and Powell, J. F. (1994). Paleobiological evidence of the peopling of the Americas: A
morphometric view. In Bonnichsen, R., and Steele, D. G. (eds.), Method and Theory for Investi-
gating the Peopling of the Americas, Center for the Study of the First Americans, Corvallis, OR,
pp. 141164.
Steen-McIntyre, V., Fryxell, R., and Malde, H. E. (1981). Geologic evidence for age of deposits at
Hueyatlaco archeological site, Valsequillo, Mexico. Quaternary Research 16: 117.
Stone, A. C., and Stoneking, M. (1998). MtDNA analysis of a prehistoric Oneota population: Implica-
tions for the peopling of the New World. American Journal of Human Genetics 62: 11531170.
Storck, P. J. (1991). Imperialists without a state: The cultural dynamics of early Paleoindian colonization
as seen from the Great Lakes region. In Bonnichsen, R., and Turnmire, K. L. (eds.), Clovis: Origins
and Adaptations, Center for the Study of the First Americans, Corvallis, OR, pp. 153162.
Storck, P. J., and Spiess, A. E. (1994). The significance of new faunal identifications attributed to an
Early Paleoindian (Gainey complex) occupation at the Udora site, Ontario, Canada. American
Antiquity 59: 121142.
Stothers, D. M. (1996). Resource procurement and band territories: A model for lower Great Lakes
Paleoindian and Early Archaic settlement systems. Archaeology of Eastern North America 24:
173216.
Stothert, K. (1985). The preceramic Las Vegas culture of coastal Ecuador. American Antiquity 50:
613637.
Struever, S., and Holton, F. A. (1979). Koster: Americans in Search of Their Past, Doubleday, New
York.
Szabo, B. J., Malde, H. E., and Irwin-Williams, C. (1969). Dilemma posed by uranium-series dates
on archaeologically significant bones from Valsequillo, Puebla, Mexico. Earth and Planetary
Science Letters 6: 237244.
Tankersley, K. B. (1998). Variation in the early Paleoindian economies of Late Pleistocene eastern
North America. American Antiquity 63: 720.
Tankersley, K. B., Tompkins, C., and Kuzmin, Y. V. (1996). Clovis precursors: The Siberian roots of
fluted points. Paper presented at 61st Annual Meeting of the Society for American Archaeology,
New Orleans.
Taylor, R. E. (1991). Frameworks for dating the Late Pleistocene peopling of the Americas. In Dillehay,
T. D., and Meltzer, D. J. (eds.), The First Americans: Search and Research, CRC Press, Boca Raton,
FL, pp. 77112.
Taylor, R. E. (1992). Radiocarbon dating of bone: To collagen and beyond. In Taylor, R. E., Long, A.,
and Kra, R. S. (eds.), Radiocarbon After Four Decades, Springer-Verlag, New York, pp. 375
402.
Taylor, R. E., Payen, L. A., Prior, C. A., Slota, P. J., Jr., Gillespie, R., Gowlett, J. A. J., Hedges, R. E. M.,
Jull, A. J. T., Zabel, T. H., Donahue, D. J., and Berger, R. (1985). Major revisions in the Pleistocene
age assignments for North American human skeletons by C-14 accelerator mass spectrometry:
None older than 11,000 C-14 years B. P. American Antiquity 50: 136140.
Taylor, R. E., Haynes, C. V., Jr., and Stuiver, M. (1996). Clovis and Folsom age estimates: Stratigraphic
context and radiocarbon calibration. Antiquity 70(269): 515525.
Taylor, R. E., Kirner, D. L., Southon, J. R., and Chatters, J. C. (1998). Letter to Science. Science
280(5367): 1171.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 99

Thompson, L. G., Yao, T., Davis, M. E., Henderson, K. A., Mosley-Thompson, E., Lin, P.-N., Beer, J.,
Synal, H.-A., Cole-Dai, J., and Bolzan, J. F. (1997). Tropical climate instability: The last glacial
cycle from a Qinghai-Tibetan ice core. Science 276: 18211825.
Thompson, R. S. (1985). The age and environment of the Mount Moriah (Lake Mohave) occupation at
Smith Creek Cave, Nevada. In Mead, J. I., and Meltzer, D. J. (eds.), Environments and Extinctions:
Man in Late Glacial North America, Center for the Study of Early Man, University of Maine,
Orono, pp. 111120.
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 continental radiation of the four founding
Native American mitochondrial DNAs. American Journal of Human Genetics 53: 563590.
Torroni, A., Neel, J. V., Barrantes, R., Schurr, T. G., and Wallace, D. C. (1994). Mitochondrial DNA
clock for the Amerinds and its implications for timing their entry into North America. Proceed-
ings of the National Academy of Sciences 91: 11581162.
Tuck, J. A., and McGhee, R. (1976). An Archaic Indian burial mound in Labrador. Scientific American
235(5): 122129.
Turner, C. G., II (1983). Dental evidence for the peopling of the Americas. In Shutler, R., Jr. (ed.),
Early Man in the New World, Sage, Beverly Hills, pp. 147158.
Turner, C. G., II (1986). Dentochronological separation estimates for Pacific rim populations. Science
232: 11401142.
Turner, C. G., II (1987). Late Pleistocene and Holocene population history of East Asia based on dental
variation. American Journal of Physical Anthropology 73: 305321.
Turner, C. G., II (1994). Relating Eurasian and Native American populations through dental morphol-
ogy. In Bonnichsen, R., and Steele, D. G. (eds.), Method and Theory for Investigating the Peopling
of the Americas, Center for the Study of the First Americans, Corvallis, OR, pp. 131140.
Tuross, N., and Dillehay, T. D. (1995). The mechanism of organic preservation at Monte Verde, Chile,
and one use of biomolecules in archaeological interpretation. Journal of Field Archaeology 22:
97110.
Ukraintseva, V. A., Agenbroad, L. D., and Mead, J. I. (1996). A palaeoenvironmental reconstruction
of the mammoth epoch of Siberia. In West, F. H. (ed.), American Beginnings: The Prehistory
and Palaeoecology of Beringia, University of Chicago Press, Chicago, pp. 129136.
Underhill, D. A., Li, J., Zemans, R., Oefner, P. J., and Cavalli-Sforza, L. L. (1996). A Pre-Columbian Y
chromosome-specific transition and its implications for human evolutionary history. Proceedings
of the National Academy of Sciences 93: 196200.
Vartanyan, S. L., Arslanov, K. A., Tertychnaya, T. V., and Chernov, S. B. (1995). Radiocarbon dating
evidence for mammoths on Wrangel Island, Arctic Ocean, until 2000 BC. Radiocarbon 37(1):
16.
Vasilievsky, R. S. (1996). Ustinovka 1. In West, F. H. (ed.), American Beginnings: The Prehistory and
Palaeoecology of Beringia, University of Chicago Press, Chicago, pp. 255267.
Waters, M. R. (1985). Early man in the New World: An evaluation of the radiocarbon-dated pre-Clovis
sites in the Americas. In Mead, J. I., and Meltzer, D. J. (eds.), Environments and Extinctions: Man
in Late Glacial North America, Center for the Study of Early Man, University of Maine, Orono,
pp. 125144.
Waters, M. R. (1986). Sulphur Springs Woman: An early human skeleton from southeastern Arizona.
American Antiquity 51: 361365.
Waters, M. R., Forman, S. L., and Pierson, J. M. (1997). Diring Yuriakh: A Lower Paleolithic site in
central Siberia. Science 275: 12811284.
Watson, E., Forster, P., Richards, M., and Bandelt, H.-J. (1997). Mitochondrial footprints of human
expansions in Africa. American Journal of Human Genetics 61: 691704.
Weber, R. L. (1993). Current research: The Amazon, eastern Brazil, and the Orinoco. American Antiq-
uity 58: 151154.
Weiss, K. M. (1994). American origins. Proceedings of the National Academy of Sciences 91: 833835.
West, F. H. (1981). The Archaeology of Beringia, Columbia University Press, New York.
West, F. H. (ed.) (1996). American Beginnings: The Prehistory and Palaeoecology of Beringia, Uni-
versity of Chicago Press, Chicago.
Whitley, D. S., and Dorn, R. I. (1993). New perspectives on the Clovis vs. pre-Clovis controversy.
American Antiquity 58: 626647.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

100 Fiedel

Whitney-Smith, E. (1998). Ecological impact of Paleoindian through second order overkill: A systems
dynamic perspective. Paper presented at 63rd annual meeting of Society for American Archaeol-
ogy, Seattle.
Willey, G. R. (1966). New World archaeology in 1965. Proceedings, American Philosophical Society
110(2): 140145, Philadelphia.
Willey, G. R. (1974). New World prehistory: 1974. American Journal of Archaeology 78: 321331.
Willig, J. A. (1991). Clovis technology and adaptation in far western North America: Regional pattern
and environmental context. In Bonnichsen, R., and Turnmire, K. L. (eds.), Clovis: Origins and
Adaptations, Center for the Study of the First Americans, Corvallis, OR, pp. 91119.
Willig, J. A. (1996). Environmental context for early human occupation in western North America. In
Akazawa, T., and Szathmary, E. J. E. (eds.), Prehistoric Mongoloid Dispersals, Oxford University
Press, Oxford, pp. 241253.
Willig, J. A., and Aikens, C. M. (1988). The Clovis-Archaic interface in far western North America. In
Willig, J. A., Aikens, C. M., and Fagan, J. L. (eds.), Early Human Occupation in Far Western North
America: The Clovis-Archaic Interface, Anthropological Papers, No. 21, Nevada State Museum,
Carson City, pp. 140.
Wilmsen, E. N., and Roberts, F. H. H., Jr. (1984). Lindenmeier, 19341974: Concluding Report on
Investigations, Smithsonian Contributions to Anthropology 24, Smithsonian Institution Press,
Washington, DC.
Wormington, H. M. (1957). Ancient Man in North America, 7th ed., Denver Museum of Natural History
Popular Series, No. 4, Denver.
Wright, H. E., Jr. (1991). Environmental conditions for Paleoindian migration. In Dillehay, T. D., and
Meltzer, D. J. (eds.), The First Americans: Search and Research, CRC Press, Boca Raton, FL,
pp. 113136.
Wobst, M. (1976). Locational relationships in Palaeolithic society. Journal of Human Evolution 5:
4958.
Yamaura, K. (1998). The sea mammal hunting cultures of the Okhotsk Sea with special reference to
Hokkaido prehistory. Arctic Anthropology 35(1): 321334.
Yesner, D. R. (1998a). Colonization models, archaeological signatures, and early sites in interior Alaska.
Paper presented at 63rd Annual Meeting of the Society for American Archaeology, Seattle.
Yesner, D. R. (1998b). Origins and development of maritime adaptations in the northwest Pacific region
of North America: A zooarchaeological perspective. Arctic Anthropology 35(1): 204222.
Yi, S., and Clark, G. (1983). Observations on the Lower Paleolithic of Northeast Asia. Current Anthro-
pology 24: 181202.
Yi, S., and Clark, G. (1985). The Dyuktai Culture and New World origins. Current Anthropology 26:
120.
Young, D., and Bettinger, R. L. (1997). Computer simulations of the colonization of continents. Paper
presented at 62nd Annual Meeting of the Society for American Archaeology, Nashville.
Young, D., Patrick, S., and Steele, D. G. (1987). An analysis of the Paleoindian double burial from
Horn Shelter no. 2, in central Texas. Plains Anthropologist 32: 257279.

BIBLIOGRAPHY OF RECENT LITERATURE

Agenbroad, L., Mead, J., and Nelson, L. (1990). Megafauna and Man: Discovery of Americas Heart-
land, Mammoth Site of Hot Springs, SD.
Akazawa, T., and Szathmary, E. J. E. (eds.) (1996). Prehistoric Mongoloid Dispersals, Oxford Univer-
sity Press, Oxford.
Aldenderfer, M. (1999). The Pleistocene-Holocene transition in Peru and its effects upon human use
of the landscape. Quaternary International 53/54: 1119.
Anderson, D. G., and Gillam, J. C. (2000). Paleoindian colonization of the New World: Implications
from an examination of physiography, artifact distributions, and demography. American Antiquity
(in press).
Ardila Calderon, G. I. (1991). The peopling of northern South America. In Bonnichsen R., and Turnmire,
K. L. (eds.), Clovis: Origins and Adaptations, Center for the Study of the First Americans,
Corvallis, OR, pp. 262282.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 101

Bard, E., Hamelin, B., Fairbanks, R. G., and Zindler, A. (1990). Calibration of the 14 C timescale over
the past 30,000 years using mass spectrometric U-Th ages from Barbados corals. Nature 345:
405410.
Boldurian, A. T. (1990). Lithic technology at the Mitchell locality at Blackwater Draw: A stratified
Folsom site in eastern New Mexico. Memoir 24, Plains Anthropologist 35(130): 1115.
Bonnichsen, R. (1991). Clovis origins. In Bonnichsen, R., and Turnmire, K. L. (eds.), Clovis: Origins
and Adaptations, Center for the Study of the First Americans, Corvallis, OR, pp. 309330.
Bonnichsen, R., and Steele, D. G. (eds.) (1994). Method and Theory for Investigating the Peopling of
the Americas, Center for the Study of the First Americans, Corvallis, OR.
Bonnichsen, R., and Turnmire, K. L. (eds.) (1991). Clovis: Origins and Adaptations, Center for the
Study of the First Americans, Corvallis, OR.
Borrero, L. (1996). The Pleistocene-Holocene transition in southern South America. In Straus, L. G.,
Eriksen, B. V., Erlandson, J. M., and Yesner, D. R. (eds.), Humans at the End of the Ice Age: The
Archaeology of the Pleistocene-Holocene Transition, Plenum Press, New York, pp. 339354.
Borrero, L. (1999). Human dispersal and climatic conditions during Late Pleistocene times in Fuego-
Patagonia. Quaternary International 53/54: 9399.
Bradley, B. A. (1993). Paleoindian flaked stone technology in the North American High Plains. In
Soffer, O., and Praslov, N. D. (eds.), From Kostenki to Clovis, Plenum Press, New York, pp. 251
262.
Brown, K. L. (1980). A brief report on Paleo-Indian-Archaic occupation in the Quiche Basin,
Guatemala. American Antiquity 45: 313324.
Carlisle, R. (ed.) (1988). Americans Before Columbus: Ice-Age Origins, Ethnology Monographs,
No. 12, Department of Anthropology, University of Pittsburgh, Pittsburgh.
Carlson, R. L. (1998). Coastal British Columbia in the light of North Pacific maritime adaptations.
Arctic Anthropology 35(1): 2335.
Collins, M. B. (1999). Clovis Blade Technology: A Comparative Study of the Keven Davis Cache,
Texas, University of Texas Press, Austin.
Crawford, M. H. (1998). The Origins of Native Americans: Evidence from Anthropological Genetics,
Cambridge University Press, Cambridge.
Curran, M. L. (1996). Paleoindian in the Northeast: The problem of dating fluted point sites. The Review
of Archaeology 17(1): 211.
Dillehay, T. D. (1991). Disease ecology and initial human migration. In Dillehay, T. D., and Meltzer,
D. J. (eds.), The First Americans: Search and Research, CRC Press, Boca Raton, FL, pp. 231
266.
Dillehay, T., Ardila Calderon, G., Politis, G., and de Moraes Courinho Beltrao, M. C. (1992). Earliest
hunters and gatherers of South America. Journal of World Prehistory 6: 145204.
Dincauze, D. F. (1996). Large Paleoindian sites in the Northeast: Pioneers marshalling camps? Bulletin
of the Massachusetts Archaeological Society 57: 317.
Dixon, E. J. (1993). The Quest for the Origins of the First Americans, University of New Mexico Press,
Albuquerque.
Easton, N. A. (1992). Mal de mer above Terra Incognita, or what ails the coastal migration theory?
Arctic Anthropology 29(2): 2842.
Edwards, R. L., Beck, J. W., Burr, G. S., Donahue, D. J., Chappell, J. M. A., Bloom, A. L., Druffel,
E. R. M., and Taylor, F. W. (1993). A large drop in atmospheric 14 C/12 C and reduced melting in
the Younger Dryas, documented with 230 Th ages of corals. Science 260: 962968.
Ellis, C., and Deller, D. B. (1997). Variability in the archaeological record of northeastern early Pale-
oindians: A view from southern Ontario. Archaeology of Eastern North America 25: 130.
Ellis, C., Goodyear, A. C., Morse, D. F., and Tankersley, K. B. (1998). Archaeology of the Pleistocene-
Holocene transition in eastern North America. Quaternary International 49/50: 151166.
Erlandson, J., and Moss, M. L. (1996). The Pleistocene-Holocene transition along the Pacific coast of
North America. In Straus, L. G., Eriksen, B. V., Erlandson, J. M., and Yesner, D. R. (eds.), Humans
at the End of the Ice Age: The Archaeology of the Pleistocene-Holocene Transition, Plenum Press,
New York, pp. 278302.
Fagan, B. (1987). The Great Journey: The Peopling of Ancient America, Thames and Hudson, London.
Fiedel, S. J. (1992). Prehistory of the Americas, 2nd ed., Cambridge University Press, Cambridge.
Frison, G. C., and Bonnichsen, R. (1996). The Pleistocene-Holocene transition on the Plains and Rocky
Mountains of North America. In Straus, L. G., Eriksen, B. V., Erlandson, J. M., and Yesner, D. R.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

102 Fiedel

(eds.), Humans at the End of the Ice Age: The Archaeology of the Pleistocene-Holocene Transition,
Plenum Press, New York, pp. 303318.
Gamble, C. (1994). Timewalkers, the Prehistory of Global Colonization, Harvard University Press,
Cambridge, MA.
Gibbons, A. (1996). The peopling of the Americas. Science 274: 3133.
Gibbons, A. (1998). Calibrating the mitochondrial clock. Science 279: 28.
Gnecco, C. (1999). An archaeological perspective of the Pleistocene-Holocene boundary in northern
South America. Quaternary International 53/54: 39.
Goodyear, A. C. (1982). The chronological position of the Dalton horizon in the southeastern United
States. American Antiquity 47: 382395.
Goodyear, A. C., III (1989). A hypothesis for the use of cryptocrystalline raw materials among Pa-
leoindian groups in North America. In Ellis, C., and Lothrop, J. C. (eds.), Eastern Paleoindian
Lithic Resource Use, Westview, Boulder, CO, pp. 19.
Goslar, T., Arnold, M., Bard, E., Kuc, T., Pazdur, M. F., Ralska-Jasiewiczowa, M., Rozanski, K.,
Tisnerat, N., Walanus, A., Wicik, B., and Wieckowski, K. (1995). High concentration of atmo-
spheric 14 C during the Younger Dryas cold episode. Nature 377: 414417.
Greenberg, J. H., Turner, C. G., II, and Zegura, S. L. (1986). The settlement of the Americas: A
comparison of the linguistic, dental, and genetic evidence. Current Anthropology 27: 477497.
Haynes, G. (1991). Mammoths, Mastodonts, and Elephants: Biology, Behavior, and the Fossil Record,
Cambridge University Press, Cambridge.
Hoganson, J. W., and Ashworth, A. C. (1992). Fossil beetle evidence for climatic change 18,000
10,000 years B.P. in south-central Chile. Quaternary Research 37: 101116.
Holliday, V. T. (1987). A reexamination of Late-Pleistocene boreal forest reconstructions for the south-
ern High Plains. Quaternary Research 28: 238244.
Holliday, V. T. (1997). Paleoindian Geoarchaeology of the Southern High Plains, University of Texas
Press, Austin.
Hughen, K. A., Overpeck, J. T., Lehman, S. J., Kashgarian, M., Southon, J., Peterson, L. C., Alley, R.,
and Sigman, D. M. (1998). Deglacial changes in ocean circulation from an extended radiocarbon
calibration. Nature 391: 6568.
Jelinek, A. J. (1992). Perspectives from the Old World on the habitation of the New. American Antiquity
57: 345347.
Kelly, R. L. (1996). Ethnographic analogy and migration to the Western Hemisphere. In Akazawa,
T., and Szathmary, E. J. E. (eds.), Prehistoric Mongoloid Dispersals, Oxford University Press,
Oxford, pp. 228233.
Kitigawa, H., and van der Plicht, J. (1998). Atmospheric radiocarbon calibration to 45,000 yr B.P.:
Late Glacial fluctuations and cosmogenic isotope production. Science 279: 11871190.
Levesque, A. J., Mayle, F. E., Walker, I. R., and Cwynar, L. C. (1993). A previously unrecognized
Late-Glacial cold event in eastern North America. Nature 361: 623626.
Levine, M. A. (1990). Accommodating age: Radiocarbon results and fluted point sites in northeastern
North America. Archaeology of Eastern North America 18: 3364.
Lotter, A. F. (1991). Absolute dating of the Late-Glacial period in Switzerland using annually laminated
sediments. Quaternary Research 35: 321330.
MacDonald, D. H. (1998). Subsistence, sex and cultural transmission in Folsom culture. Journal of
Anthropological Archaeology 17: 217239.
MacDonald, G. F. (1968). Debert, a Paleo-Indian site in central Nova Scotia. Anthropology Papers,
No. 16, National Museums of Canada, Ottawa.
Markgraf, V. (1991). Younger Dryas in southern South America? Boreas 20: 6369.
Martin, P. S., Thompson, R. S., and Long, A. (1985). Shasta ground sloth extinction: A test of the
blitzkrieg model. In Mead, J. I., and Meltzer, D. J. (eds.), Environments and Extinctions: Man
in Late Glacial North America, Center for the Study of Early Man, University of Maine, Orono,
pp. 514.
Maschner, H. D. G. (1997). American Beginnings and the archaeological record of Beringia: A
comment on variability. Antiquity 71(271): 223228.
McAvoy, J. M., and McAvoy, L. D. (1997). Archaeological Investigations of Site 44SX202, Cactus
Hill, Sussex County, Virginia, Research Report Series, No. 8, Virginia Department of Historic
Resources, Richmond.
P1: FMN
Journal of Archaeological Research [jar] PL120-74 August 30, 1956 22:3 Style file version Nov. 19th, 1999

The Peopling of the New World 103

Mead, J. I., and Meltzer, D. J. (eds.) (1985). Environments and Extinctions: Man in Late Glacial North
America, Center for the Study of Early Man, University of Maine, Orono.
Meltzer, D. J. (1993). Pleistocene peopling of the Americas. Evolutionary Anthropology 1(5): 157169.
Meltzer, D. J., and Mead, J. I. (1985). Dating Late Pleistocene extinctions: Theoretical issues, analytical
bias, and substantive results. In Mead, J. I., and Meltzer, D. J. (eds.), Environments and Extinctions:
Man in Late Glacial North America, Center for the Study of Early Man, University of Maine,
Orono, pp. 145174.
Miotti, L. (1992). Paleoindian occupations at Piedra Museo locality, Santa Cruz province, Argentina.
Current Research in the Pleistocene 9: 3031.
Mock, C. J., and Bartlein, P. J. (1995). Spatial variability of Late-Quaternary paleoclimates in the
western United States. Quaternary Research 44: 425433.
Morse, D. F. (1996). The Sloan Site: A Paleo-Indian Dalton Period Cemetery in the Western Lowlands
of Northeast Arkansas, Smithsonian Institution Press, Washington, DC.
Nichols, J. (1990). Linguistic diversity and the first settlement of the New World. Language 66: 475
521.
Preston, D. (1995). The mystery of Sandia Cave. The New Yorker, June 12, pp. 6683.
Saunders, J. J. (1992). Blackwater Draws: Mammoths and mammoth hunters in the terminal Pleistocene.
In Fox, J. W., Smith, C. B., and Wilkins, K. T. (eds.), Proboscidean and Paleoindian Interactions,
Baylor University Press, Waco, TX, pp. 123148.
Shott, M. J. (1992). Radiocarbon dating as a probabilistic technique: The Childers site and Late
Woodland occupation in the Ohio Valley. American Antiquity 57: 202230.
Shutler, R., Jr. (1985). Dating the peopling of North America. In Mead, J. I., and Meltzer, D. J. (eds.),
Environments and Extinctions: Man in Late Glacial North America, Center for the Study of Early
Man, University of Maine, Orono, pp. 121124.
Stanford, D., and Day, J. (eds.) (1992). Ice Age Hunters of the Rockies, University of Colorado Press,
Niwot.
Steele, J., Adams, K., and Sluckin, T. (1998). Modelling Paleoindian dispersals. World Archaeology
30: 286305.
Stepp, D. (1997). Paleobiology focuses on first Americans: Genetic clues point to one early migration.
Mammoth Trumpet 12(3): 1, 911.
Tankersley, K. B. (1994). Was Clovis a colonizing population in eastern North America? In Dancey,
W. S. (ed.), The First Discovery of America, Archaeological Evidence of the Early Inhabitants of
the Ohio Area, Ohio Archaeological Council, Columbus, pp. 95116.
Webb, E., and Rindos, D. (1997). The mode and tempo of the initial human colonization of empty
landmasses: Sahul and the Americas compared. In Barton, C. M., and Clark, G. A. (eds.), Re-
discovering Darwin: Evolutionary Theory in Archeological Explanation, Archeological Papers,
No. 7, American Anthropological Association, Washington, DC, pp. 233250.
Wittkofski, J. M., and Reinhart, T. R. (eds.) (1989). Paleoindian Research in Virginia: A Synthesis,
Special Publication 19, Archeological Society of Virginia, Richmond.
Yesner, D. R. (1998). Origins and development of maritime adaptations in the northwest Pacific region
of North America: A zooarchaeological perspective. Arctic Anthropology 35(1): 204222.

Você também pode gostar