You are on page 1of 45

REVIEW ARTICLE

The Millennium before Clovis


Gary Haynes
University of Nevada – Reno, NV
Note text corrections on pages 137, 150, & 160 that were not in the final printed version
This paper reviews the published information, uncertainties about claims, and possible technological and
cultural relationships of a sample of sites which have older-than-Clovis dates in North America. The goal is
to trace the origins of “Classic” Clovis techno-cultural patterns. Some sites in the sample contain lithic
artifacts and some do not. Production technology and artifact characteristics in a number of the lithic sites
(such as Debra Friedkin and possibly Page-Ladson) may be evidence of Clovis ancestry, but the lithic
materials in most pre-Clovis sites cannot be explicitly linked to Clovis. A few nonlithic sites (such as
Manis, Firelands, and Lindsay) may indicate a pre-Clovis pattern of large-mammal exploitation
foreshadowing a later Clovis trait. Overall, the available data are incomplete or ambiguous, and as a
result, individual interpretations have produced incompatible models of Clovis origins.
Keywords older-than-Clovis, pre-Clovis, proto-Clovis, Clovis

1. Introduction and organization (before 1950 CE/AD). Radiocarbon dates needing


This paper examines a sample of prehistoric North calibration were entered in the Calib 7.0 online
American sites that predate the Clovis era by 1000 program, which uses the IntCal13 calibration curve
14
C years or so (Table 1), and evaluates their possible (Reimer et al. 2013). Calibrated ages are generally
evidence about Clovis origins. One of the dominant given here as approximate midpoints in the 2-sigma
unanswered questions in prehistory is whether Clovis range, unless the range is very wide. The notation
was a replacement or a descendent culture, or, as “ka” means 1000 years before Y2K (2000 CE/AD).
Kunz (2010, 144) put it, whether the people were The term Clovis is used here to designate a horizon
“immigrants or home grown.” of traits left by human groups who made fluted
The search for Clovis origins goes back to the bifaces across almost all of unglaciated North
1960s and 1970s; C. Vance Haynes (1964, 1966, 1967, America around 13,000 cal yr BP. Waters and
1969, Stafford (2007) set the upper and lower bounds of
1970, 1974, 1980, 1982, 1987, 1991a, 1993), now Clovis at 11,050
retired, was perhaps the best-known scientist engaged and 10,800 14C yr BP (tentatively calibrated with
in the search. It became clear to him and others maximum span at 13,250–12,800 cal yr BP, and at
helping to define Clovis that the distinctive technologi- minimum span 13,125–12,925 cal yr BP). A later
cal traits, extensive mobility, and subsistence pub- lication (Waters and Stafford 2013, 550) sets a
strategies did not originate in North America, but slightly
from a north- ern Eurasian Upper Paleolithic different “maximum period for Clovis – 13,100–
background. It was also obvious that there had to be 12,600 cal yr BP,” and another publication
an ancestral human population in the Americas, (Rasmussen et al. 2014, which includes Waters and
which developed the Classic Clovis features. Haynes Stafford in its author list) cites the minimum age at
12,556–12,707 cal yr BP, stretching the range later to
and others had faith that a pre-Clovis progenitor
encompass the relatively young direct date (10,705 ±
population would be found. It seems logical to expect
35 14C yr BP, or ca. 12,556–12,707 cal yr BP) on the
that sites dating to the millennium before Clovis
Anzick (MT) site’s Clovis-associated child.
would provide evidence about the origins of Clovis.
The narrow time range has been met with question-
ing (Haynes et al. 2007), because older ages have been
1.1 Timeframe published, such as an average of two radiocarbon
Radiocarbon dates and calibrations are cited as orig- dates from the Aubrey site, 11,565 14C yr BP (13,400
inally reported, respectively 14C yr BP and cal yr BP cal yr BP) (Ferring 2001, 49, table 3.2), which
Correspondence to: gahaynes@unr.edu
Waters and Stafford (2013, 544) consider possibly con-
taminated or not directly associated with the Clovis
134
© 2015 W. S. Maney & Son Ltd
and the Center for the Study of the First Americans
DOI 10.1179/2055556315Z.00000000016 PaleoAmerica 2015 VOL. 1 NO. 2
Haynes The Millennium before Clovis

Table 1
Sites reviewed in this paper, showing radiocarbon and calibrated
ages
Direct dating of organic artifacts or
Site Approximate age of materials bones?

Monte Verde II, Chile 12,780–12,230 14C yr BP Yes, on possible organic artifacts and
seaweed
Meadowcroft Rockshelter, 11,300–12,800 14C yr BP in middle of Stratum II (Miller Yes, on bark-like organic fragment
PA point minimum age 14,000 cal yr BP); 12,800–16,175 14C
yr BP in lower Stratum II; and 19,600 14C yr BP date on
bark-
like-fragment
Cactus Hill, VA 10,920 14C yr BP on Clovis level charcoal; 15,070 14C yr No, on charcoal flecks
BP
on charcoal below Clovis; OSL ages 17–20 ka No, only OSL sediment estimates, with
Topper, SC OSL estimate of Clovis 13,500 ± 1,000 cal yr BP; OSL some radiocarbon dates on
ages below Clovis >15,200 ± 1,500 cal yr BP; and questionable hearth
deeply buried “hearth” >50,000 14C yr BP
Buttermilk Creek Complex OSL estimates of lower part of Clovis level 14,350 ± 910 No, only OSL dates on sand grains
(Debra L. Friedkin cal yr BP and 14,070 ± 910 cal yr BP; OSL estimates of within fluvial sediments
site), TX BCC 14,080 ± 920 to 16,515 ± 1,075 cal yr BP
Gault, TX Preliminary OSL age estimates similar to BCC No
Swan Point, AK 12,360–11,770 14C yr BP on residues, 12,060 14C yr BP on Yes, on various organics
ivory 12,040 14C yr BP on charcoal
Broken Mammoth, AK 11,770 14C yr BP on charcoal No, on charcoal
Mead, AK 11,600 and 11,560 14C yr BP on charcoal No, on
charcoal
Big Eddy, MO “Early/Middle Paleoindian” (including Clovis) level: No, on charcoal
10,260–11,384 14C yr BP; pre-Clovis: 4400–12,950 14C yr
BP
Paisley Caves, OR 12,265 and 12,165 14C yr BP on oldest coprolites having Yes, on coprolites
human mtDNA
Hebior and Schaefer, WI 12,290–12,590 14C yr BP on mammoth bone collagen Yes, on marked bones
Burning Tree, OH 11,660 and 11,450 14C yr BP on presumed mastodont gut Yes, on presumed gut contents and
contents; 10,860 14C yr BP on mastodont bone collagen bone
Coats-Hines, TN 12,050–12,030 14C yr BP on charred material just above No, on charcoal
artifacts and just below mastodont bones
Lindsay, MT 9,490–11,925 14C yr BP on unpurified mammoth bone Yes, on bone collagen
collagen; 12,105–12,330 14C yr BP on purified collagen;
and 12,270–12,300 14C yr BP on purified collagen
Page-Ladson, FL Average of six dates 12,450 14C yr BP (14,475 cal yr BP) No, organics in sediment were dated
Manis, WA 11,990 14C yr BP on mastodont rib Yes, on mastodont bone
Ayer Pond, WA 11,700–11,990 14C yr BP Yes, on possibly impacted,
chopped bison bone
Firelands, OH 11,740 14C yr BP (13,550 cal yr BP) Yes, on bone

archeology, and thus not validly dating the Clovis photheres are associated with Clovis points in
component; and a terminus post quem date from the Sonora, Mexico (Sanchez et al. 2013, 2014),
East Wenatchee site, 11,600 ± 50 14C yr BP (also and another date
roughly 13,400 cal yr BP) (Kuehn et al. 2009),
which
Waters and Stafford (2013, 543) observe is only a
maximum limiting date, not necessarily the age of
the Clovis materials. Miller et al. (2013) accept
Ferring’s (2012) argument that the 11,565 14C yr BP
age at Aubrey is securely associated with the fluted-
point occupation, which Ferring terms “proto-
Clovis,” because it lacks some of the distinctive
(“Classic”) Clovis features of overshot flaking and
large-blade manufacturing, as defined for Clovis by
Bradley et al. (2010).
Two recently acquired radiometric dates now offer
additional possible support for a much earlier start
of the Clovis era. One date (11,560 ± 140 14C yr BP;
roughly 13,450 cal yr BP) on wood charcoal comes
from a repeatedly occupied kill- and camp-site, El
Fin del Mundo, where the bones of at least two gom-
PaleoAmerica 2015 VOL. 1 NO. 2 13
Haynes The Millennium before Clovis
(11,626 ± 68 14C yr BP, roughly 13,400 cal yr BP) comes short of the opposite edge. Furthermore, Huckell
from a piece of wood charcoal excavated near an (2014, 151) has pointed out that it is not certain if
overshot flake at the Beach biface cache site in intentional overshot flaking was “uniquely diagnostic
North Dakota, which Huckell (2014, 151) has argued is of Clovis.” This means that although the presence of
Clovis, although it lacks fluted points. overshot flaking may still be potentially diagnostic of
Regarding the overshot flaking on Clovis bifaces, Eren Clovis (Eren et al. 2014, 60), by itself and without
et al. (2013, 2014) strongly argued that overshot flaking other features such as fluted bifaces or macroblades,
of bifaces – the removal of flakes from one it cannot be considered decisively diagnostic.
edge that cross the entire face and take off part of the In this paper, the start date for Clovis (meaning
opposite edge – was not an intentional Clovis strategy; both the Classic sensu Ferring (2012) and Miller
instead it was a mistake, and overface or “ultrashot” et al.
flaking was the intention, with flakes terminating just

(2013) and pre-Classic, which is Ferring’s “proto- available about the dated materials and how they
Clovis”) is set at 11,600 14C yr BP (∼13,400–13,495 spatially relate to the artifacts. However, the
cal yr BP), while the cryptic existence of what I had soundness of age assignments may be open to re-
earlier termed proto-Clovis (Haynes 2002, 253), evaluation, along with presump- tions about human
meaning the stage before even fluting had been behavior associated with different objects.
invented, might be detectable within the millennium
before Clovis. My review focuses on a sample of 1.3 Background: Anatomically modern
sites with ages back to 12,600 ± 200 14C yr BP Homo sapiens arrives in the Americas
(roughly 14,100–15,500 cal yr BP at 2-sigma) Goebel et al. (2008) summarized knowledge about
(Table 1). I mention older claims when they may be the nature and tempo of modern human entry into
relevant to understanding archeological developments the Americas. Ancient mitochondrial DNA
in the Americas. (mtDNA) evidence shows that a European-derived
The possibility that people co-existed with Clovis population
but used very different lithic technological methods
and subsistence practices bears on the question of
Clovis origins. For some time, there has been debate
about whether the Clovis horizon marks the appear-
ance of North America’s earliest typable projectile
point. One possible contemporary is Goshen (Frison
1996; Sellet et al. 2009), and another is Western
Stemmed (Beck and Jones 1997, 2010; Bryan 1988;
Bryan and Tuohy 1999; Fiedel and Morrow 2012;
Goebel and Keene 2014; Willig and Aikens 1988).
However, the evidence for contemporaneity of
stemmed points and Clovis is inconclusive, or
negative in the case of Goshen (Waters and Stafford
2014). If human groups with different production
technologies co-existed at the time of Clovis, often
using the same geographic ranges, their potential
cultural, social, and genetic relationships with each
other will be difficult to model.

1.2 Minimally acceptable evidence


for archeological claims
Most sites examined here have been excavated under
controlled conditions, although their original discov-
ery may have been accidental. Their stratigraphy has
been formally described, but spatial associations may
be ambiguous in some cases. In most sites, the
artifacts are unquestionably of human manufacture,
but in a few sites, the materials are not solidly proven
to have been made or modified by human hands.
Some form of dating has been done, not always
directly on arti- facts, and adequate details are
13 PaleoAmerica 2015 VOL. 1 NO. 2
Haynes The Millennium before Clovis
of Homo sapiens was in Northeast Asia by 19,880 ± corridor livable as much as 2000 14C years earlier (ca.
160 14C yr BP (roughly 24,000 cal yr BP; Raghavan 14,000–15,000 cal yr BP or before) based on
et al. 2013), and later must have become admixed radiocarbon-dated wood macrofossils and optically
with East Asian populations, eventually evolving the stimulated luminescence (OSL)-dated postglacial
types of mtDNA found in Recent Native Americans dunes in the corridor (Munyikwa et al. 2011; other
(e.g., Rasmussen et al. 2014). By 22 ka, humans had references in Ives et al. 2013, 151–152).
occupied parts of western Beringia. Indigenous The Pacific coastal route has been prominently
Northeast Asians, Recent Native Americans, and a touted as a more likely way southward from
sample of human skeletons from North America’s Beringia, because (1) biface shapes from Siberia,
late Pleistocene are clearly connected genetically. Japan, Oregon, California’s Channel Islands, and
Native American ancestors originated exclusively South America are perceived to be morphologically
from Asian populations, with early European admix- similar enough to imply a shared origin (Erlandson
ture in Siberia long before dispersing to the Americas. 2013); (2) terrestrial fauna such as bear and caribou
Some time after 17 ka, human groups began disper- occupied isolated refugia along the British Columbia
sing out of Beringia into continental North and South coast ca. 16,000 cal yr BP; (3) the discovery of
America. The route taken is still anyone’s guess. Both interti- dal latest Pleistocene/early Holocene sites has
the iced-over coasts and interior landscapes of North led to optimism that now submerged late-glacial sites
America had been bottlenecks or barriers until around in the
16–14 ka (Clague et al. 2004; Munyikwa et al. Pacific Northwest region of Canada eventually may
2011). Ives et al. (2013, 162) proposed that “in be found (Mackie et al. 2013); and (4) some sites
ecological terms human populations could have reen- located around “the Western margins of the
tered [sic] the [ice-free] Corridor region as early as American continents” (Mackie et al. 2013, 145)
11,600 14C BP [roughly 13,400 cal BP],” which is a appear to date relatively early.
time estimate that places availability of the inland
route at the same age as first dates on Clovis a thou- 2. Finds, claims, and possibilities
sand miles south of the corridor. This conservative Here I first summarize information about a sample of
age estimate for the corridor is based only on mammal- archeological (or potentially archeological) sites
bone dates – there are other estimates that make the dating >1000 14C years older than Clovis, then
discuss another sample of lithic-bearing sites with
ages nearer or within ∼1000 14C years of Clovis, and spanning from >50,000 years ago through
complete the review by examining a sample of Clovis and the Holocene. This site yielded bones
animal-bone sites without lithics that date to within of numerous
∼1000 14C years of Clovis. While the discussions Pleistocene and Holocene animals, including
sum- marize issues that I consider important and some interpreted as tools, a pendant, and by-
unresolved products of marrow extraction, and apparent
about the sites, I also accept that the sites provide human finger prints or palm prints in heated clay
hints about cultural developments before the Clovis that are up to 37,000 14C years old (Chrisman
era. The examinations are centered on North et al. 1996, 2003; but also see Schaffer and
America. Latin American sites dating to the Baker 1997). The senior investigator R.
millennium before Clovis are only briefly MacNeish died just before the publi- cation of a
mentioned. large book about the site, and his absence may
The samples were taken from others’ lists (such as in explain why his provocative findings are not
Collins et al. 2013; Waters and Stafford 2013), and are very actively discussed.
not exhaustive. I have not included plenty of possible
sites such as those listed in Faught and Freeman 2.1 Sites dating ≥1000 14C years before
(1998), because most are inadequately published or CORRECTION
Clovis "..broken in
the claims made about them have been vigorously
Several archeological sites have claims of"antiquity,wit
questioned. A prominent site not discussed is Sandia h LGM humans
exceptionally early evidence about humanpossibly
Cave (Haynes and Agogino 1986; Hibben 1941;
behavior in the Americas, dating to before theresponsible,
although he
Stevens and Agogino 1975), with its Sandia type 2
last glacial maximum (LGM) considers
basally thinned point, once discussed as a possible that
(23–18 ka). In eastern North America, lithicunlikely."
Clovis point precursor. The Sandia type may be as
sites with very early dates ranging to 27,240 ± 230
old as Clovis, older than Clovis, or younger than 14
C yr BP (roughly 31,000 cal yr BP) have been
Clovis (Haynes and Agogino 1986) – and the site reported on
remains in the limbo of uncertainty. Another site not the Delmarva peninsula: Parsons Island, Miles
included here is Pendejo Cave, which MacNeish Point, Oyster Cove, and Cators Cove (Lowery et
(2004) proposed had a record of human occupation
PaleoAmerica 2015 VOL. 1 NO. 2 13
Haynes The Millennium before Clovis
al. 2010; Stanford and Bradley 2012). Another site,
down, because of the extreme separation in time (at
Cinmar, is located 65 km offshore and 74 m below
least 7000 years for the youngest site, and more than
sea surface in the Atlantic; a bipointed biface
twice that for others), and the prospect that any
was pulled up
shared technological features such as overshot
there in the same dredge with mastodont remains
flaking reflect cultural convergence rather than a
dated 22,760 ± 90 14C yr BP (roughly 27,000 cal yr
lineal relationship with Clovis. Perhaps as important
BP). The early Delmarva lithics are generally well-
an issue is the dating of these pre-LGM claims,
made bifaces and other kinds of tools. Any possible
which open them to question. They are much older
ancestral connection to Clovis cannot be nailed
than models in archeology (summarized in Goebel
et al. 2008; also see Graf 2008, 2009) and genetics
(e.g., Mulligan and Kitchen 2013; Perego et al. 2009)
establish for the first human movement out of
Northeast Asia into the Americas. The controversial
attribution of this possible complex to Solutreans
from Iberia must confront the chronological problem
that the Solutrean lithic industry developed from the
Gravettian ca. 20,000 14C yr BP, almost 3000 14C
years after the Cinmar mastodont died.
Holen and Holen (2013) proposed that seven mid-
continental sites are pre-LGM and seven sites date to
the LGM, created by mammoth-steppe-adapted
humans who had moved east from Asia between
40,000 and 22,000 14C yr BP. Two examples of the
sites are Lovewell I, Kansas, dated 20,430 ± 300 14C
yr BP (about 24,000 cal yr BP), and Villa Grove,
Colorado, dated 33,405 ± 340 14C yr BP (about
37,000 cal yr BP). Both are without lithics and
contain broken mammoth bones interpreted (Holen
and Holen 2013, and other references therein) as
impacted by humans using stone percussors. A third
example is Cooperton, Oklahoma (Anderson 1962,
1975; Mehl 1975), another lithics-free mammoth find
with claims for bone-breakage by humans, dated
17,575 ± 550–20,400 ± 450 14C yr BP (roughly
21,000–25,000 cal yr BP). A fourth example not in
the Holen and Holen (2013) list is the lithics-free
mammoth site called Inglewood, Maryland, with a
bone date of 20,070 ± 265 14C yr BP (roughly 24,000
cal yr BP) and a date on associated woody vegetation
of 29,650 ± 750 14C yr BP (roughly 33,000 cal yr BP).
The find was first interpreted (Haynes 1991, 236,
figure 6.11) as a partial skeleton with many axial
bones in anatomical order but with larger elements
( particularly appendicular and cranial) broken by
heavy equipment while the bones were still enclosed
in waterlogged sediments. Karr (2015) has attempted
to re-interpret it as an example of bones that were
humanly flaked in antiquity, but the conclusions are
insupportable and based on incomplete information.
Until the site has been adequately published, it
cannot be considered further here.
With one exception, Burnham, Oklahoma, dated
36,000–35,000 14C yr BP (about 40,000 cal yr BP),
the 14 sites listed in Holen and Holen (2013) and
the
Inglewood site contained either no lithics or no more

13 PaleoAmerica 2015 VOL. 1 NO. 2


Haynes The Millennium before Clovis
than four flaked-stone items, mostly flakes. The
scarcity or lack of lithic technological information may not be convinced that these specimens were
and the lengthy time separation of these sites from actually made by humans.
Clovis preclude reliable attempts to trace
lineal/cultural
relationships.
An example of a pre-LGM site in South America is
Arroyo del Vizcaino, Uruguay, with a possibly butch-
ered extinct giant sloth (Fariña et al. 2014) dated
27,000–30,000 14C yr BP (about 31,000–34,000 cal
yr BP) and possible lithics, including flakes and a poss-
ible scraper with micropolish that was found “in very
close association with several bones” (Fariña et al.
2014, 5, figure 2). Even a vague link to Clovis is
hard to see with this site, both technologically and
chronologically. For one thing, Clovis itself has not
been found in Uruguay, although undated “Clovis-
like fluted points” are known from as far south as
southern Chile, on the other side of South America
(Jackson 2006, 116, figure 6.4, 117). The sloth site
and the Clovis-like horizons in South America are
floating too far apart in time (and space) to hint at a
possible cultural connection.
Examples of South American claims with pre-
LGM ages and definable lithic technological features
are Vale de Pedra Furada (not the same as
Boqueirão de Pedra Furada) and Toca da Tira
Peia, both in Brazil, among others (e.g., see Boëda
et al. 2013; Lahaye et al. 2013). These are not well
known and not widely accepted by North
American archeologists, probably because (1) the
lithic assemblages are much cruder looking than
those made by modern Homo sapiens living at the
same time in Europe, Asia, and Africa, which is an
argument Boëda et al. (2013) reject as unscientific;
(2) the sites are very rare and scattered in space,
thus not appearing to be archeological cultures on
the landscape, another bit of reasoning rejected by
Boëda et al. (2013, 462) who claim the earliest
human populations were very small, “dispersed
across vast areas,” and “without apparent links
between them;” and (3) the dates are much older
than current genetic and archeological models
accept for human migrations out of northern Asia,
which is also a line of reasoning that Boëda et al.
(2013) and some other South American prehistorians
see as naively overlooking the extremely patchy
nature of the first migrants who would have disap-
peared without archeological or genetic trace.
The lithics in these Brazilian sites, OSL dated to
>20 ka, are forms such as asymmetrically and irregu-
larly flaked cobbles and “beaked” flakes (see the
numerous figures in Boëda et al. 2013). They are not
at all similar to what have been defined as Clovis
tech- nological characteristics (Bradley et al. 2010),
and cannot be linked to Clovis. Many archeologists
PaleoAmerica 2015 VOL. 1 NO. 2 13
Haynes The Millennium before Clovis
Other South American pre-Clovis lithic sites (e.g., definitely shaped lithic tools, broken and unbroken
those listed in Collins et al. 2013; Dillehay 2013; pebbles of a certain size range that may have been
Waters and Stafford 2013) are not so distant in time selected from naturally occurring pebbles at the site,
from the Clovis era. According to Waters and Stafford and a variety of organic materials such as seaweed,
(2013, 545) at least five sites have ages that bone fragments, and plant fibers. Radiocarbon dates
fall within the Clovis era, while a “handful” have on wood identified as artifacts and on seaweed from
older-than-Clovis ages. Lavallée (2000 [orig. 1993]: the lowest cultural component cluster between
42) on one page counted “about 10 sites” in South 12,780 ± 240 and 12,230 ± 140 14C yr BP (roughly
America with dates older than Clovis, while on 15,000–14,000 cal yr BP), which are 1100 to >800
14
another page (Lavallée 2000 [orig. 1993]: 51) she C years older than Clovis.
counted “a dozen or so.” Numerous other lists of South Questions or issues? A number of professionals have
American sites possibly older than Clovis or the found aspects of Monte Verde II singular, peculiar, or
same age can be found elsewhere (e.g., Aceituno et al. confusingly documented (e.g., Fiedel 1999; Lynch
2013; Dillehay 1999; Kelly 2003; MacNeish 2001; West 1996, 1999), even after a first round of
2004). The best known of such sites surely must be criticisms was aggressively addressed (Dillehay et al.
Monte Verde, Chile. 1999; see Fiedel n.d., for response to the response),
or they have questioned the correctness of age assign-
ments in the oldest component (e.g., Dickinson 2011;
2.1.1 MONTE VERDE II, CHILE Haynes 1999). Some are not convinced that all the
Monte Verde II, Chile (Dillehay 1989, 1997; Dillehay lithic and organic materials found at Monte Verde II
et al. 2008) is undoubtedly the most extensively are actually artifacts (e.g., Haynes 1999; West 1999).
described and publicized South American site older A 2012 survey of opinions from over 200 individuals
than Clovis. Another locus nearby, called Monte Verde who were actively involved in publishing or
I, has lithic objects and hearth-like features that gathering data about the peopling of the
may date to ca. 33,000 14C yr BP (about 37,000 cal yr Americas (Wheat
BP), but it has not been adequately explored or 2012) found that 33 per cent of respondents either
published. Monte Verde II has features interpreted as rejected or were undecided about Monte Verde’s
hearths and hut remains, wooden artifacts, a few claims as a pre-Clovis site. Most of the respondents
were academic archeologists, along with a number of
genetic anthropologists, skeletal biologists, linguists, many being annual updates, but no full
and a few others. publication exists that provides all the updated
Relation to Clovis? If the age assignments and information in one place and answers to
interpretations are correct, Monte Verde II and other critiques. Examples of sum- maries (and
sites listed by Waters and Stafford (2013) and Collins defensive responses to criticisms) are
et al. (2013) show that at the time of Clovis and Adovasio et al. (1977, 1978, 1999), while more
proto-Clovis in North America, and possibly for details about specific aspects of the site’s earliest
several centuries before, South American people had com- ponents are in Fitzgibbon et al.
very distinct material cultures, even in the same (1982) and
regions where Clovis-like patterns would be, or Stuckenrath et al. (1982).
already were, manifest, such as fluted points, associ- Stratum IIa, one of 11 natural strata identified
ation with extinct megafauna, and Clovis-era in the rockshelter, has radiocarbon dates on
radiocar- bon dating (Jackson 2006; Jackson et al. wood charcoal
of 11,300 ± 700 to 12,800 ± 870 14C yr BP (setting
a possible range of 11,250–17,500 cal yr BP) in its
middle part, which contained a resharpened
unfluted
lanceolate biface (typed as a Miller point; Figure
1) similar to an incomplete biface found at the
relatively nearby Krajacic open-air site, and
what was called a

2007; Ranere and Cooke 1991).


Figure 1 Miller type lanceolate biface from
Meadowcroft Rockshelter Stratum IIa, as drawn by Joel
2.1.2 MEADOWCROFT ROCKSHELTER, PENNSYLVANIA
Gunn (edited from a figure in Adovasio et al. 1977,
Another important site is Meadowcroft Rockshelter, figure 24). This specimen has been used in publications
Pennsylvania. There are abundant short publications to illustrate the Miller type (e.g., Stanford and Bradley
about this North American site from 1975 onward, 2012, figure 4.12), but in fact it is a

14 PaleoAmerica 2015 VOL. 1 NO. 2


Haynes The Millennium before Clovis
reworked/resharpened piece whose dimensions and shape
are not as originally manufactured. The 1977 description
possible fragment of a fluted point (Fitzgibbon et al.
(Adovasio et al. 1977, 46–47) stated its outline had been 1982, 96), which seems to have been recognized later
“altered by breaking and re-shaping the point.” A scar from a as unrelated to Clovis. A lower part of the stratum
probable “substantial impact” is visible on the right hand con- tained several hundred flaked-stone artifacts
view; the point was subsequently reshaped by pressure and has
flaking. The edges of the base were ground up to the lateral
tick marks on the figure.
dates ranging from 12,800 ± 870 to 16,175 ± 975 14C
yr BP (calibrating to a range of roughly 13,000–
22,000 cal yr BP). Two fragmentary human bones
were also found in a firepit in lower Stratum IIa
(Sciulli 1982), assigned dates of 13,270 ± 340 14C yr
BP and 13,240 ± 1010 14C yr BP (calibrating roughly
between 13,200 and 18,300 cal yr BP). A carbonized
fragment of bark-like material from the deepest part
of the stratum, said to be cut and consistent in mor-
phology with basketry plaiting strips (or warp or
weft fragments) (Stuckenrath et al. 1982, 83, table 3;
see also Adovasio and Pedler 2013), was dated
19,600 ± 2400 14C yr BP (roughly 17,900–28,500 cal
yr BP); note the very large sigma and the very wide
calibration range.
Noteworthy is the fact that the most distinctive fea-
tures of Clovis technology are not represented at the
rockshelter, although fluted-point assemblages have
been found nearby in Pennsylvania and Ohio.
Adovasio et al. (1999, 418; the passage is also
repeated verbatim in Adovasio and Pedler 2005, 26)
called the
assemblage part of a “technologically standardized
and sophisticated, small, polyhedral core- and blade-
based industry of decidedly Eurasiatic, Upper
Paleolithic ‘flavor’” [quotation marks in orig.], which
is “precisely the sort of lithic reduction strategy that
should be evidenced at this time” [italics in original].
Questions or issues? Some vocal skeptics, such as
Haynes (1991b) and Mead (1980), were not convinced
that the site’s oldest dates were correct because of
important inconsistencies – for example, only
Holocene species of flora and fauna were identified
in the sediments dated to the Pleistocene, and there
was no noticeable stratigraphic break at the
Pleistocene–Holocene boundary. However, Adovasio
et al. (1999; Adovasio and Pedler 2005) retorted
that
both of these characteristics are common in sites
from nearby areas and elsewhere. Another questioner,
Kelly (1987, 332), wondered if bioturbation could have
mixed charcoal and artifacts of different ages, or led
to contamination of radiocarbon samples. Below the
cul- tural assemblage in Stratum IIa is much older
carbo- nized material (in Stratum I), but no cultural
items, which may have been a source of
contamination. Stratum IIa has radiocarbon dates
stretching over 13,000 years, a very slow
sedimentation rate, and
fairly thick sterile layers in it, indicating “an
enormous amount of time […] encompassed in a

PaleoAmerica 2015 VOL. 1 NO. 2 14


Haynes The Millennium before Clovis
relatively thin deposit which accumulated very slowly” (Flannery 2003, 52) who reviewed Adovasio’s and
(Kelly 1987, 332). A respected Australian paleoecologist Page’s (2002) mass-market book The First Americans
suggested that anomalous flora and fauna, the exca- G of McAvoy and McAvoy 1997). Points of
vators’ miscommunication about the site’s radiocar- recognized post-Clovis types were found below
bon samples, and a lack of professionalism when Clovis, and their
responding to skeptics were major stumbles (his
term) that made the rockshelter’s reported chronology
suspicious.
The 2012 survey of opinions from individuals
involved in researching the peopling of the Americas
(Wheat 2012) found that over 60 per cent of respon-
dents either rejected the claims about Meadowcroft
or were undecided — in other words, not convinced.
Relation to Clovis? The imprecise dating is trou-
bling; still, some materials seem to be older than
Clovis. They lack the most diagnostic lithic features
that would link them less ambiguously to Clovis,
such as ultrashot or overface flaking on bifaces.
However, the early levels did yield lithic materials
superficially consistent with what has been found in
some fluted-point assemblages, such as blades and lan-
ceolate bifaces. The Miller lanceolate biface (Figure
1) from a pre-Clovis level is similar enough to be
called by some archeologists a precursor to the
Clovis point, although this piece had been reshaped
after breaking; Adovasio (1993, figure 7) illustrated a
poss- ible prototype of the Miller form, based on the
Meadowcroft and Krajacic finds, neither of which
was in pristine condition. In Meadowcroft publi-
cations, pre-Clovis lithic materials from the site are
compared favorably to Clovis lithics, and to both pre-
Clovis and post-Clovis assemblages from the
eastern and western United States. However, Collins
and Lohse (2004, 182) state that “neither the points
nor the blades are technologically very similar to
those of Clovis.” The blades and the Miller lanceolate
biface are also similar to post-Clovis materials from
the eastern USA – for example, some variants of a
late Paleoindian/early Archaic point type called Hi-
Lo in the Great Lakes region are difficult to dis-
tinguish from the Miller lanceolate (Ellis 2004, 71),
and blades continued to be made by post-Clovis cul-
tures of the Great Lakes region and the Ohio and
Tennessee valleys, although with reduced emphasis
(Ellis 2004; Kimball 1996; Sherwood et al. 2004).

2.1.3 CACTUS HILL, VIRGINIA


Unlike Meadowcroft, the Cactus Hill site did yield
fluted points, along with other lithic materials –
early Archaic through Woodland projectile point
types above Clovis, and blades and small bifaces
occurring 7–15 cm below the Clovis level. The site
has an unconsolidated medium-fine sand matrix with
acknowledged uncertainties about context and strati-
graphic integrity (McAvoy and McAvoy 1997, 179;
see also Johnson 1997, unnumbered page in appendix

14 PaleoAmerica 2015 VOL. 1 NO. 2


Haynes The Millennium before Clovis
positions are attributed to drift. Charcoal from the between the Clovis layer and the underlying pre-
lower part of a Clovis hearth was dated 10,920 ± 250 14C Clovis stratum is very lengthy, yet there is no
yr BP (about 12,500 cal yr BP at the midpoint of erosional disconformity or sign of weathering,
the 2-sigma range). The upper part of the hearth also which is a geoarcheological problem, and the
contained “partly carbonized” Holocene-age wood temporal separ- ation may be even longer if the OSL
(McAvoy and McAvoy 1997, 167). A date on “an dates are accepted. In the Wheat (2012) survey of
amorphous scatter of carbon” associated with quartzite professional opinions about sites older than Clovis,
lithics below the Clovis hearth was 15,070 ± 70 14C more than 75 per cent
yr BP (roughly 18,200 cal yr BP). The suite of OSL were unconvinced about Cactus Hill.
age estimates on four sediment Relation to Clovis? The two oldest projectile points
samples from below the fluted-point level, ranging from Cactus Hill have lanceolate and semi-triangular
from 17 to 20 ka, “if anything […] are apt to be too forms (Figure 2), somewhat similar to the
young,” according to Feathers et al. (2006, 185).
Questions or issues? Archaic type projectile points
also occur in the Clovis-bearing level in some of the
site (Units 1/9 and 2/9) (McAvoy and McAvoy
1997, table 15b, unnumbered page in appendix A),
apparently in pits dug into the Clovis level from above.
Charcoal dates that are both Holocene and late
Pleistocene from a Clovis hearth might call into
question the site’s stratigraphic integrity. However,
Feathers et al. (2006, 185) think the “archeology, sedi-
mentology, pedology, botany, radiocarbon, and lumi-
nescence” lines of evidence are in agreement, and the
site strata have “overall integrity” (Feathers et al.
2006, 167), in spite of deflation, turbation, low accre-
Figure 2 One of the two subtriangular bifaces found below
tion rates, and slight mixing of deposits. the Clovis level at Cactus Hill (edited from an illustration in
The 4000 14C year (or almost 6000 cal year) gap Stanford and Bradley 2012, figure 4.12; drawn by Marcia
Bakry; used by permission of D. Stanford).

Meadowcroft Miller type but otherwise not especially human-made fire feature, dated >50,000 14C yr
similar technologically to Clovis bifaces. The same BP, which is very near the early limit for
goes for the blades. Superficially, the pre-Clovis radiocarbon
lithics are not what is typical in a Clovis assemblage dating and cannot be calibrated with IntCal13.
(Collins and Lohse 2004, 182), and they are also very “Unusual chert artifacts” (Goodyear 2005, 109)
distantly separated in time from Clovis, by such as broken cobbles without negative bulbs
4000–6000 calendar years, making their possible and frequent
ancestral relationship very uncertain. The pre-Clovis- hinge terminations, and small-flake clusters, are
dated assemblage, like the Meadowcroft pre-Clovis said to be cultural. The lithics, which do not
assemblage, may or may not be a contender for change in character between levels dated
ances- tor to Clovis or proto-Clovis (sensu Haynes >50,000 and 15,000
14
2002, not Ferring 2012). C yr BP, are characterized as a smashed core
and microlithic industry (Goodyear 2005),
2.1.4 TOPPER, SOUTH CAROLINA matching nothing else in Paleoindian prehistory
Topper is a stratified, multicomponent site containing or in any hypothetical Old World ancestral
a quarry-related Clovis occupation (Miller 2010; region. Noteworthy is the fact that the broken
Smallwood et al. 2013) in colluvial-fluvial sediments. chert objects derive from bedrock outcrops of
The pre-Clovis claims have not been well published the chert, now buried at the site. Questions or
in refereed journals or a monograph. Goodyear issues? Besides the abundant artifacts, Topper
(2005) proposed that the site also has lithics in a has yielded thousands of local chert fragments,
paleo- sol that are 3000–7000 calendar years older most of which are not thought to be humanly
than Clovis, as estimated by OSL dating and pro- duced. Waters et al. (2009, 1309) suggested
reckoning of that chert
the time needed for the paleosol to have developed breakage may have been caused by thousands of
in alluvial sand. A precise age is not known. years of natural fires, freeze–thaw cycles, or
Goodyear also interpreted carbonized plant remains stream transport. The very old specimens with
found well below the Clovis levels as remnants of a characteristics

PaleoAmerica 2015 VOL. 1 NO. 2 14


Haynes The Millennium before Clovis
of human-made lithics may be the result of noncul-
tural fragmentation. Natural breakage events must
have occurred thousands of times in Topper’s last
50,000 years; so many natural events could have
produced a proportion of broken specimens that
look like artifacts, which is not extraordinary at all
but in fact almost inevitable, according to statistical Figure 3 Examples of small broken chert pieces found at
principles (e.g., see Hand 2014). This is similar to Topper, from sediments dated by OSL as 3000–7000
calendar years older than Clovis (edited from an online
the reasoning used by Meltzer, Adovasio, and
photograph accessed 1 September 2014,
http://www.allendale- expedition.net/museum/bb2.jpg).

Dillehay (Meltzer et al. 1994) in doubting that


humans flaked the stones in the lower levels at
Brazil’s 35–50 ka Boqueirão de Pedra Furada, and
also similar to the logic of Haynes (1973) who
argued that natural events created the flaked-stone
specimens at California’s 200 ka Calico Early Man
Archaeological Site.
In the Wheat (2012) survey of professional opinions,
85 per cent of respondents were unconvinced about
Topper’s pre-Clovis claims. Some recent reviews of
sites older than Clovis (e.g. Adovasio and Pedler
2013; Collins et al. 2013; Waters and Stafford 2013)
either do not mention or do not give serious attention
to Topper.
Relation to Clovis? The mostly small lithics below the
Clovis level (Figure 3) are not widely thought to be cul-
turally produced. OSL age estimates have large error
ranges for the lithics said to be older than Clovis,
and the fact that the materials are technologically
unchanged for over 35,000 calendar years seems
unusual in the late Quaternary prehistory of Homo
sapiens. The technology of these materials, if they are
artifacts and not nonculturally broken pieces, is very
different from Classic Clovis, hence making any
possible relationship with Clovis indeterminable. The
very oldest date also does not fit with current genetic
models of human entry into the Americas.

2.1.5 DEBRA L. FRIEDKIN (BUTTERMILK CREEK COMPLEX), TEXAS


The Buttermilk Creek Complex (BCC) in the open-
air Debra L. Friedkin site is a lithics-only assemblage
below a level which did not contain fluted points but
did have channel flakes and evidence of overshot
flaking, leading to classification as Clovis. The
Clovis level underlies a series of time-diagnostic
artifacts from late Prehistoric at the top through
Archaic, late Paleoindian, and Folsom. The age
assignment of the BCC is 15.5–13.2 ka, based on 18
OSL dates on quartz grains in clay-rich overbank
deposits (Waters et al. 2011a). Most of the BCC
assemblage is debitage, but also found were bifaces,
blade fragments, mostly small-flake tools, often with
use-wear on edges, and small blades (Waters et al.
2011a, supporting online material, p. 27).
14 PaleoAmerica 2015 VOL. 1 NO. 2
Haynes The Millennium before Clovis
Questions or issues? Morrow et al. (2012) point out are not completely identical to other Clovis
that the OSL dates are imprecise and do not provide
a decisive age for the pre-Clovis materials; the stan-
dard deviations for the dates are huge, some over
1000 calendar years, effectively establishing age
ranges of 2000 calendar years for each sample. Also,
some OSL ages from the Folsom and Clovis strati-
graphic zones are demonstrably too old, by about
1000 calendar years, suggesting that OSL
overestima- tion of the BCC age is also possible.
Morrow et al. (2012) further speculate that younger
debitage from levels above the BCC may have moved
downward into the BCC level, perhaps due to the
opening of ver- tical cracks during shrinkage of the
clay-rich matrix, technically a vertisol, or possibly
through animal (e.g., crayfish) burrowing, although
Driese et al. (2013) concluded that the site sediments
do not contain evidence for significant mixing of
cultural hor- izons (also see Jennings and Waters
2014, 44, note 3 for other supporting references).
Relation to Clovis? Jennings and Waters (2014) com-
pared technological and typological traits from three
assemblages, the Friedkin site’s Clovis materials, the
Gault (see below) Clovis assemblage, and the BCC.
The Clovis assemblages at Friedkin and Gault were
significantly different from each other in some
respects, such as frequencies of large debitage and
overshot flakes and relative counts of debitage types,
while the BCC differed from both Clovis
assemblages in some ways. But the BCC materials
had 6 of 10 traits
often used to “define Clovis” (Jennings and Waters
2014, 38, table 12). The BCC lacked fluted points,
bifaces with overshot flaking, blade cores, and
retouched blades. For the most part, these distinctions
were interpreted as significant and meaningful and
not the result of sample-size differences.
Morrow et al. (2012, 3680) point out that the BCC’s
reported “end-thinning flakes, partial overshot flakes
[…], an overshot flake […], biface reduction debris,
[…] edge-modified flakes, utilized blades and blade-
lets, graver, notches, bend-break tools, and tools
made on radially broken pieces […] also occur in
Clovis assemblages.” Nevertheless, a few BCC traits
are uncommon in Clovis assemblages from other
sites, such as the presence of burin spalls, discoidal
core reduction, and deliberate radial breakage of
bifaces (Jennings and Waters 2014, 39). The differ-
ences may mean the BCC cannot be considered a
form or variant of Clovis, but on the other hand, the
similarities may indicate that the BCC is technologi-
cally related to Clovis, perhaps a precursor lacking
only the production trait of fluting.
The analysis presented in Jennings and Waters
(2014) supports the interpretation of the BCC assem-
blage as a Clovis progenitor, because the materials

PaleoAmerica 2015 VOL. 1 NO. 2 14


Haynes The Millennium before Clovis
assemblages and the age is older than Clovis; but Questions or issues? Morrow et al. (2012) raised
Jennings and Waters also cautiously point out that all some questions about Gault. According to Collins
Clovis assemblages are not identical to each other (2007) and Waters et al. (2011b), who examined the
anyway, for a variety of possible reasons. On the other Gault stratigraphy, the site has evidence for disturb-
hand, Morrow et al. (2012) reason that the BCC ance and mixing of objects of different ages, mostly
assemblage could be part of a Clovis occupation, and in upper ( post-Clovis) levels. There is possible tram-
is not necessarily a progenitor because the pre- Clovis pling damage on lithic artifacts in the Clovis zone,
dating is so imprecise. which could have caused downward movement of
some lithics. Depositional rates were “modest”
2.1.6 GAULT, TEXAS before, during, and after Clovis times (Collins 2007,
The Gault site is 250 m upstream from the Debra 61). Root penetration and microbioturbation pro-
L. Friedkin site on Buttermilk Creek. Both sites are cesses have occurred, such as insect burrowing,
near high-quality lithic toolstone outcrops and reliable which could have mixed materials vertically. Parts of
springs, and both contain abundant lithics, mostly some refitted lithic pieces had been vertically separated
debitage. “Preliminary OSL dates” (Collins et al. 2013, by 19 cm. Further contributing to uncertainty about
528) have placed the lowermost Gault lithics at about the site’s depositional integrity is the presence below
the same age as the Debra L. Friedkin site’s the Clovis zone of the very small stemmed points
pre-Clovis materials, up to 2000 (or more) calendar similar in morphology to later Holocene projectile
years older than Clovis. Lithic technological differ- point types. Collins et al. (2013) suggest that these
ences from Clovis are highlighted by Collins et al. and a slightly larger similar piece from the Debra
(2013) and Jennings and Waters (2014), such as in L. Friedkin site are first glimpses of a Pleistocene
biface morphology and manufacturing techniques. The tech- nology so far unknown from Paleoindian
unique presence of very small projectile points (Collins contexts.
et al. 2013, 531, figure 30.7) also differs from Clovis. Relation to Clovis? As with the BCC, the oldest
The site does have technological charac- teristics materials at Gault are not precisely dated, although
shared with Clovis such as prismatic blades and they are found under Clovis-specific levels. They
similarities of point preforms. A 2 × 2 m stone pavement may be very early ( proto?) Clovis, or considerably
is thought to be older than Clovis, with chert flakes on older.
one side of it and large-animal bones on another.
The existence of the very small points in the older- 2.2.1 ALASKAN FINDS
than-Clovis material is especially surprising, since Several culture-historical components have been
these forms are not Paleoindian-related and do not named in east Beringia, but there are
appear anywhere else in North America until much disagreements about labeling in different
later in the Holocene. Prismatic blades in the oldest classification systems (Goebel and Buvit 2011;
assemblages are similar to Clovis forms, but not all Potter 2011). The East Beringian Tradition
bifaces are similar to those in Clovis assemblages (Holmes 2001, 2011; also see West 1981, 1996)
from other sites. The pre-Clovis materials plausibly is also called simply the Beringian Tradition
could be technological precursors to Clovis, but they (Hoffecker 2005; Hoffecker and Elias 2007), and
do seem distinct in some ways from the Clovis what is called Nenana by some (Goebel
assem- blage (Jennings and Waters 2014). Yet it is et al. 1991; Hoffecker et al. 1993; Vasil’ev 2011)
important to keep in mind that Clovis assemblages is pre-
from many North American sites do have variability, ferentially termed Diuktai by others (Holmes
and the differences among them probably reflect 2001), while the Denali complex may also be
differences in site function, location, availability of termed American Paleoarctic or even Diuktai.
lithic raw materials, occupation times, and other The relative ages of these complexes are not
factors. Pre-Clovis differences from Clovis at Gault clear; a figure in
or at the Debra L. Friedkin site may be a result of Dixon (2013, 122, figure 6.8) shows
variations in these same kinds of factors, and are not Denali/eastern
necessarily Beringian as oldest, followed by Nenana above,
outright indications of a major cultural/technological Denali again above that, northern Paleoindian,
dichotomy. If such factors are better studied, the tech- Denali again, northern Archaic, and late Denali.
nological and typological relationships between The oldest sites share dominant features such
Clovis and vaguely similar older-than-Clovis traits as presence of microblades and burins. Examples
may become clearer (Jennings and Waters 2014). of important sites are Swan Point, Dry Creek,
and Broken Mammoth, among a few others that
2.2 Sites dating ≤1000 14C years before Clovis date before the Clovis era (see Potter et al.
14 PaleoAmerica 2015 VOL. 1 NO. 2
Haynes The Millennium before Clovis
2013, 83, table
Dates from some of the Alaskan sites’ earliest cul-
5.1 for a list). Rasic (2011) reports a site that, while
tural components partly fall within my espoused
it yielded mostly Sluiceway type stemmed points,
Clovis time range. For example, at the Mead site, the
also had one fluted point, one lanceolate point, and
lowest cultural component has charcoal dates of
a single microblade core (with a date of 11,200 ± 40
14
11,560 ± 80 and 11,600 ± 60 14C yr BP
C yr BP (about 13,100 cal yr BP), which is within
(about
the Clovis era). There are others with similar ages 13,400 cal yr BP), which Waters and Stafford (2007)
just before the Younger Dryas. would consider older than Clovis, but which here are
considered to be coeval with the earliest Clovis dates.
At Broken Mammoth, the oldest component has char-
coal dates of 11,770 ± 210 and 11,770 ± 220 14C yr BP
(about 13,800 cal yr BP), placing it 170–700 14C
years (or an average of about 200 cal years)
older than
Clovis, depending on which of the assigned Clovis
time spans is preferred. Swan Point’s oldest dates are
on various materials, including ivory collagen
(12,060 ± 60 14C yr BP, or about 13,900 cal yr BP),
charcoal (12,040 ± 40 14C yr BP, or about 13,850 cal
yr BP), and lithic-surface residues (12,360 ± 60,
12,220 ± 40, 12,110 ± 50, 12,100 ± 40, and 11,770
±
140 14C yr BP, calibrating to a maximum of roughly
14,800 cal yr BP). These ages are older than Clovis
by 170–760 14C years (or, when calibrated, range
from approximately coeval to 1000 cal years older).
Faunal remains include bones of game birds, water-
fowl, elk, bison, and fish, and fragments of horse
teeth and mammoth teeth and tusk.
Potter et al. (2013, 95) inferred that the original
colonization of eastern Beringia was “primarily
inland and terrestrially oriented,” and that subsistence
focused on large mammals. They also (Potter et al.
2013, 93) suggested that the faunal record of eastern
Beringia is consistent with Kelly’s and Todd’s (1988)
behavioral model of early Paleoindian foragers being
highly mobile and reliant on large animals, although
foraging also was opportunistic for other food
resources when necessary. During resource scarcity
periods, the main strategy would have been to move
into new territory. When Beringian populations
expanded south, they did so through the interior
corri-
dor, “innovating fluting technology and discontinuing
microblades (but not macroblade and burin technol-
ogy)” ( parentheses in original) (Potter et al. 2013,
95). The dispersing populations kept other lithic
traits, such as standardized end scrapers and side
scrapers, macroblades, and gravers (Goebel et al.
1991) as they moved into and through the lower 48
states (see Goebel et al. 2013).
Relation to Clovis? Lithics in the East Beringian
Tradition sites are variable, often dominated by
micro- blades and microcores, but assemblages are
similar in some ways to those of Clovis (Figure 4)
(Goebel 2004; Goebel et al. 1991). The microblade
technology clearly
PaleoAmerica 2015 VOL. 1 NO. 2 14
Haynes The Millennium before Clovis
sets them apart from Clovis – this trait is not part of
the Classic Clovis package as defined by Bradley et al.
(2010). Fluting technology appeared in Alaska no earlier
than about 10,370 14C yr BP (about

14 PaleoAmerica 2015 VOL. 1 NO. 2


Figure 4 Comparison of lithic classes from Clovis and Nenana sites (edited from an illustration in Goebel 2004, figure 11.10,
used by permission of T. Goebel).

12,000–12,400 cal yr BP), shown by multiple dates 2.2.2 BIG EDDY, MISSOURI
from the Serpentine Hot Springs site (Goebel et al. The open-air Big Eddy site has cultural components
2013; Young and Gilbert-Young 2007) and Raven that include, from the top down, Mississippian,
Bluff site (Hedman 2010). This is more than 1000 Woodland, Archaic, late Paleoindian, early/middle
14
C years after Classic Clovis had first appeared in Paleoindian (including fluted bifaces), and pre-Clovis
the lower 48 states, perhaps indicating a northward (Lopinot et al. 1998, 2000; Lopinot and Ray 2000;
“return” migration of descendant cultures 2000 14C Ray et al. 1998). The site has nine accelerator (AMS)
years after the initial entry of ancestral people into dates on its “early/middle Paleoindian” component
Beringia. Most Alaskan fluted points are multiply (containing fluted bifaces and debitage) that are
fluted, and are morphologically similar to points that within the span of the Clovis era. Below the early/
seem to post-date Classic Clovis in New England middle Paleoindian materials, three flakes and three
and the Great Basin. large cobbles were found, and farther below was a
Other lanceolate biface industries such as Mesa gravel layer containing numerous chert flakes, a large
and Sluiceway also appeared in the interior of and heavy stone that may have been used as an
Alaska, termed the northern Paleoindian (Hoffecker anvil, and a large cobble that may have been used as
and Elias 2007), at around 13,250 cal yr BP, which is a hammerstone. Charcoal fragments were “scattered
nearly the same age as Clovis in the lower 48 states. throughout the early deposits” (Lopinot and Ray
These industries are quite different from those of the 2000, 3). Sixteen AMS dates were run on wood
microblade-dominated complexes. It is not clear if char-
the northern Paleoindian industries developed in coal from the component thought to be pre-Clovis;
Alaska or moved there from lower latitudes (Smith three samples returned Holocene ages, and three
et al. 2013), as fluted points may have done. others fell within the Clovis era as defined in this
paper, although they came from below the fluted-
point level. The other 10 dates from the layer below but it may also be over 1000 14C years younger.
Clovis range from 11,910 ± 440 to 12,950 ± 120 Another (“possibly oldest”; Jenkins et a l.
14
C yr BP (approximate midpoint of calibrated 2012b, 16)
range is stemmed projectile point recovered in situ was
14,400 cal yr BP). dated
Questions or issues? The discordant charcoal dates to 11,500 ± 30 14C yr BP (about 13,300 cal yr
could be a warning sign of inconspicuous stratigraphic BP)
mixing of materials in sediments. As for the older-
than-Clovis stone items, two experts in artifact use-
wear analysis did not consider them to be unambigu-
ously artifactual (Ahler 2000; Kay 2000), but a third
examiner thought the materials possibly could be
human-modified (Dillehay 2000). The site investi-
gators (Lopinot and Ray 2007) experimented with
trampling by elephant and bison to determine if the
pre-Clovis gravel layer’s lithic flakes and modified
pebbles and cobbles could have been naturally
created, which they conceded was likely for most
flakes, but they also continued to maintain that the
possible anvilstone and cobble hammer were
humanly modified.
Relation to Clovis? The possible anvilstone, the
pos- sible hammerstone, and the flakes in the gravel
layer may be Clovis age or more than 1000 14C
years older. If older than Clovis (Ray et al. 2000),
which most AMS dates suggest, there is no direct
technologi-
cal connection between the site’s simple lithic assem-
blage and the typical materials in a Clovis site.

2.2.3 PAISLEY CAVES, OREGON


The Paisley Caves consist of several rockshelters
(Gilbert et al. 2008; Jenkins 2007; Jenkins et al.
2012a, 2012b, 2013) that have produced a number
of different claims: (1) its Western Stemmed points
are as old as or older than the Waters and Stafford
(2007) age range of Clovis; (2) some coprolite speci-
mens with Native American types of mtDNA in
them, found in the deepest deposits without
associated artifacts, are older than Clovis; and (3)
animal bones were cut by stone tools 1300 14C years
before Clovis. Questions or issues? The dates
ascribed to stemmed points at Paisley Caves are
among the very few Clovis age or pre-Clovis dates
associated with stemmed points in the
Intermountain West, out of many dozens of
published dates from numerous other localities
(Beck and Jones 1997; Fiedel and Morrow 2012;
Goebel and Keene 2014).
Importantly, three of the four ages assigned to the
oldest stemmed points at Paisley Caves are younger
than or the same age as Clovis. One stemmed point
that was not recorded in situ has estimated bracketing
dates of 12,269 ± 60 to 10,050 ± 50 14C yr BP, so it
conceivably may be 800 14C years older than Clovis,
by association with a woody twig 40 cm away and lying haplogroups – or any human mtDNA at all, except
below the level of the point, which might indicate that the for a probable contaminant – in eight samples which
point was deposited after the twig. two other labs at Copenhagen and the University of
The obsidian hydration age estimates on some lithics York had said contained the Native American
are older than Clovis, but they are also older than the mtDNA haplogroups (Jenkins et al. 2012b, 26);
radiocarbon ages in the same stratigraphic positions as the however, canid mtDNA was identified in the sample
oldest stemmed point fragments (Gilbert et al. 2008; by the Washington State lab, as the Copenhagen lab
Pinson 2010). Obsidian hydration age estimates are rough had done earlier (Gilbert et al. 2008). The
at best and may be inaccurate (Anovitz et al. 1999; Duke discrepancy
2011), because even slight differences in effective was thought to be due to “(a) differences in DNA
hydration temperature may produce errors of several survival across specimens, (b) the possibility that
thousand years for early these coprolites are nonhuman; or (c) […] the ratio
Holocene/late Pleistocene artifacts (Duke 2011, 46, 54– of contaminant to endogenous DNA was too great
58). It is also doubtful if obsidian hydration for endogenous mtDNA detection using [the
results can validly make age separations at a finer tem- Washington State University] techniques” (Jenkins
poral scale than the gross major cultural periods et al. 2012b, 27), possibly implying that control for
(Paleoindian, Archaic, etc.) (Duke 2011). The question of contamination at the Washington State University
whether stemmed points are as old as or older than Clovis lab was inferior to that at the other labs. Recently,
at Paisley Caves is unsettled. one “human” coprolite with the oldest dates
As for the coprolites, human and large canid feces are (12,400 ± 60 and 12,275 ± 55 14C yr BP)
often visually indistinguishable, and the presence of the from
canid mtDNA in the same coprolites with human Paisley Caves was identified as decidedly not human,
mtDNA, the herbivore-like micromorphology of sampled because the fecal biomarkers in it and its micromor-
coprolites (Goldberg et al. 2009; Poinar et al. 2009), and phology are diagnostic of a herbivore, not a human,
the geographically anomalous human mtDNA haplotype and the presence of human mtDNA was “probably
in two coprolites (Fiedel 2014) has caused concern about derived from an undetermined contamination
possible contami- nation or misidentification. A lab at pathway” (Sistiaga et al. 2014, 815), ironically imply-
Washington State University could not find the Native ing that the Copenhagen lab’s controls for contami-
American nation were imperfect.

It is relevant when considering this give and take Hockett and Jenkins used a checklist of features (V-
about the Paisley Caves coprolites to remember an shape, non-sinuous length, shoulders present, mul-
anecdote in Matisoo-Smith and Horsburgh (2012, tiple straight and parallel marking, etc.) to classify
61), who recount how Viking samples for a DNA cutmarks, but these may or may not be present in a
analysis were collected under both the strictest possible significant proportion of marks made by various
conditions to prevent contamination and under much agents other than stone tools, according to exper-
looser, less than ideal conditions. All workers who imental work by Dominguez-Rodrigo et al. (2010,
could have come into contact with the samples were 2012) and Krasinski (2010). For example, 30 per
tested for their own DNA. The sample collected cent of trample marks may be straight rather than
under the looser conditions returned DNA from mul- sinuous, and 10 per cent of stone-tool cuts may be
tiple individuals, but it matched none of the research- sinuous rather than straight (Krasinski 2010). Stone-
ers who could have inadvertently contaminated the tool cuts may have either a V-shaped or a U-shaped
sample. Sediment from the sample’s find spot also (flat) profile in section. The experimental tool-
did not contain any of those DNA sequences. One cutting of fresh bone for comparative purposes
could conclude that even though the Paisley Caves described by Hockett and Jenkins (2013) was
excavators’ own DNA did not match the done on defleshed pig bones, apparently not in order
Copenhagen lab’s identification of a Native to dismember or fillet meat, but specifically to make
American DNA variety, some other avenue of con- marks. Such marks are not identical to marks
tamination could have existed during the brief first inflicted while butchering a fresh carcass. As it
exposure of the coprolites. stands now, the similarities of the experimental and
A third issue is the assertion by Hockett and fossil marks are evidence that can support the
Jenkins (2013, 766) that there are stone-tool cutmarks cutting of the bones after they were defleshed, but
on pre-Clovis bones, although the identifications are need not support human actions involved in proces-
only “likely” (Hockett and Jenkins 2013, 766), sing a fresh carcass. Defleshed bones could have
based on specimens having “several of the character- been cut when trampled against sharp-edged stone
istics” of experimentally created cutmarks and not fragments in the rockshelter (see Dominguez-
having the features typical of trampled bones. Rodrigo et al. 2012).
In the Wheat (2012) survey of professional
50 per cent of respondents were unconvinced about
opinions about pre-Clovis possibilities in North
the Paisley Caves claims.
America, over
Relation to Clovis? If the Paisley coprolites are
human, and the lithics are indeed as old as or older
than Clovis, the implications are complex. The pro-
duction technology of stemmed bifaces and Clovis
differs, which Bradley (1993) thought indicated differ-
ent ancestry, although there are some similarities (see
Amick 2004 for an analysis of a Great Basin
stemmed biface cache). It seems inefficient and unli-
kely that a single cultural group would have manufac-
tured important lithic tools using two distinct
knapping strategies; therefore, two different lithic tra-
ditions seem to have existed in close proximity in the
American West for several centuries. Judging from
where the diagnostic lithic implements of both tech-
nologies have been found, such as around pluvial
lakes (Willig and Aikens 1988), some of the same
ranges were used by the two cultures, some of the
same resources may have been sought, and the
people surely must have come face to face for several
centuries, which would raise the issue of how as
niche-sharing cultures they dealt with competitive
exclusion (Fiedel 2014).

2.2.4 HEBIOR AND SCHAEFER, WISCONSIN


Hebior and Schaefer are two open-air sites in south-
eastern Wisconsin dated about a millennium older
than Clovis (Joyce 2006, 2013; Overstreet 1993,
1996; Overstreet et al. 1993, 1995; Overstreet and
Kolb 2003; Overstreet and Stafford 1997). They are
both mammoth bone beds encased in pond deposits,
and having very few stone implements within the
beds. The lithic items are unusually crude looking
(images are in Joyce 2013, 475, figures 27.6, 27.7)
when compared to Paleoindian (and especially
Clovis) lithics. Other sites often mentioned together
with them, Mud Lake and Fenske, Wisconsin, are
not discussed here because few bones survive from
the original discoveries by ditch-diggers many
decades ago. As well, a sample of the Mud Lake
ulna’s plentiful surface marks that had been classified
as stone-tool cuts by Johnson (2006, 2007) and
Joyce
(2013) could not be identified as tool cuts in another
study (Krasinksi 2010).
At Schaefer, 10 mammoth bones have over two
dozen marks interpreted as cuts or wedge marks,
and at Hebior, nine bones have marks interpreted as
produced by human actions. Cuts or chop marks
may be created on modern proboscidean carcasses
during meat removal or skeletal disarticulation
(Haynes 1991; Haynes and Krasinski 2010), but they
tend to be rare or nonexistent when expert butchers
strip meat.
An undated lithic complex called Chesrow has
been mentioned as possibly associated with Hebior and
Schaefer (e.g., Overstreet 1993). In reviewing this Figure 5 Lithics from the Hebior site, edited from a
sug- gestion, Joyce (2013, 475) inaccurately claimed photograph in Joyce (2013).

I pre- sented the Chesrow–mammoth association


“as fact”
and “a given” (Joyce 2013, 478), when clearly the
message in my discussion (Haynes 2002, 49–50) was
about the ambiguity in the dating of Chesrow and
the unproven nature of any suggested association
with pre-Clovis mammoths. Joyce (2013, 478) unequiv-
ocally stated that a link between the Chesrow
Complex
and the Hebior and Schaefer mammoths is “nonexis-
tent,” but then went on to accept it as a “working
hypothesis.” Ellis (2004) did not accept the Chesrow
association with the mammoths, and considered the
Chesrow point forms to be coeval with Hi-Lo and
Dalton materials, which post-date the Clovis era.
Chesrow’s association with Hebior and Schaefer is a
very weak possibility, which was my position in
Haynes (2002), and lacks enough support to discuss
further.
Questions or issues? It is very unlikely that the mam-
moths in these two sites were butchered by the rela-
tively small lithics found with the bones, including
irregularly flaked bifaces (Figure 5). The presence of
carnivore tooth marks and trampling marks, the
close similarities of some marks to metal-tool dings,
and the claim that pry or wedge marks have been
identified, when such an impracticable and dysfunc-
tional action of wedging or prying bones apart
would not be part of elephant-carcass processing
when handheld tools are used, leads me to question
the identification of the cutmarks.
Doubt also arises when the stratigraphy is
examined for Hebior (Figure 6). The stratigraphic
profile con- tains horizontal layers, vertical columns,
and broken offset bands of intrusive or injected
sand, called
dikes or lenses that formed “very late in the develop-
ment of the stratigraphic sequence” (Overstreet and
Kolb 2003, 7). The possibility exists that these were
routes through which much younger artifacts were
moved downward by water action into the bone bed
from above. Some small objects could have been
incorporated in the intrusive vertical and lateral sand are very different from anything expected in Clovis
bands, moving short distances in the sand and there- fore sites with large-mammal bones, or in any Upper
are not visibly abraded. Vertical and lateral dis- Paleolithic butchery site in Eurasia. There is no tech-
placement of items is a possibility in any site where nological link at all to Clovis, but proboscidean
sediments are disturbed, even if only slightly. hunting or scavenging at Hebior and Schaefer 1000
Relation to Clovis? Whether or not the sites have been
correctly interpreted as mammoth-butchering locales, the
fact that single proboscideans occur in association with Figure 6 Hebior stratigraphy. Upper photograph: view of the
mammoth bones in situ, with a cleaned profile in the
lithics is indeed reminiscent of Clovis sites. Mammoth-
background, showing lighter colored lenses and stringers
associated Clovis sites, on the other hand, do not have cutting through the sediments overlying the bone level
such heavily marked bone surfaces (although poor (photographed by Vance Holliday, used by permission).
preservation could have removed such traces from some Lower photograph: closer view of part of one cleaned wall
Clovis finds), and of course, they consistently date 1000 profile after the bones had been removed (photographed by
14 G. Haynes in 2005).
C years younger. Importantly, the lithics at both sites
14
C years before Clovis is a possible sign of a frozen bog, a difficult-to-prove possibility proposed
foraging pattern similar to Clovis, namely human by Fisher (Fisher 1984a, 1996; Fisher et al. 1994;
opportunistic targeting of increasingly vulnerable or Lepper et al. 1991).
stressed mega- fauna during the climatic oscillations Questions or issues? The bones were partially
of the late glacial. sub- merged during part of the excavation, and it is
possible that marks were created on the bone
2.2.5 BURNING TREE, OHIO
At Burning Tree, three bone clusters from a
mastodont skeleton missing one rear limb’s bones,
most of the tail, most toes, and a few other elements
were found within a peat bog in the first stages of
excavation to make a water hazard on a golf course
(Lepper et al.
1991). Radiocarbon dates were run on spruce
branches in spatial association (12,620 ± 90 and
11,720 ± 110 14C yr BP; about 14,000 cal yr BP at
midpoint of both ages), peat/soil (12,230 ± 70 14C yr
BP; about 14,400 cal yr BP), non-coniferous twigs
and organic
matter from “presumed gut contents” (11,660 ± 120
and 11,450 ± 70 14C yr BP; about 13,700–13,100 cal
yr BP at midpoint of range), and bone collagen
(10,860 ± 70 14C BP) (Lepper et al. 1991, 124). The
gut content ages are about 400–600 14C years older
than the Waters and Stafford (2007) beginning date
for Clovis, but are very near or younger than the less
conservative date preferred here (see Miller et al.
2013; Sanchez et al. 2013, 2014), placing this site
within the Clovis era. The younger date on the bone
is explained as typical of bone collagen dates
(Lepper et al. 1991, 123, citing other sources).
The find lacked lithics, but had some possible tool-
made cuts and gouges on a toe bone and ribs, along
with “dragmarks” claimed to have been made by the
bones being pulled across sand or grit, which are par- surfaces during recovery actions. Some well-
ticle sizes apparently not present in the bog where the preserved fossil bones when wet have surfaces that
bones came to rest (Fisher et al. 1991, 1994), although are easier to mark than when they harden after
grit was found on the internal surface of unfused drying. The bone surface marks should be
epiphyses. Carnivore gnaw marks were also rigorously analyzed by taphonomic experts using a
identified on some pieces. The site was interpreted as contextualized approach (Dominguez-Rodrigo et al.
a partly butchered carcass that had been transported 2010). The original bones are currently in a
away from a kill site and stored in the water of a Japanese museum, to which they were sold in 1993
for $600,000. As with Hebior, Schaefer, Manis (see
equipment such as earth-movers or draglines may
below), and several other probos- cidean finds not
have driven over and deformed the waterlogged sedi-
examined in this paper, heavy
ments enclosing the bones, abrading or fragmenting
them, before exposing the bone assemblages to discov-
ery (see Haynes 1991, 236, figure 6.1), and it is also
possible that abrasion from being forced through
sedi- ment or across other bones, or the impact of
hydraulic buckets, may be responsible for at least
some bone modifications.
Relation to Clovis? The site may have been made
by early or proto-Clovis people, based on the dating
of the presumed gut contents. If it actually is a
butchered and cached proboscidean, it makes more
credible the development of a rational late-glacial
subsistence strat- egy of opportunistically targeting
large mammals. But because there are no lithics in
the site it is unlike the proven Clovis sites with large-
mammal remains; there- fore, it cannot be clearly
identified as either a Clovis site or made by people
immediately ancestral to Clovis.

2.2.6 COATS-HINES, TENNESSEE


At the Coats-Hines site, “Charred material […] from
the top of the artifact-bearing deposit” in a paleo-
pond (Deter-Wolf et al. 2011, 152) was dated
12,050 ± 60 14C yr BP (ca. 13,750–14,000 cal yr BP
at 2 sigma), about 400 14C years older than the
Clovis start date preferred here. A younger
radiocar- bon date came from just above a
mastodont bone, while a date of 12,030 ± 40 14C yr
BP came from beneath another mastodont bone, and
an older date
of 14,750 ± 220 14C yr BP came from below yet
another mastodont bone. Deeper bone-bearing depos-
its at the site were dated around 23–28,000 14C yr BP
(roughly 27,300–33,500 cal yr BP). Mastodont bones
are said to have been marked by stone-tool edges,
although adequate illustrations of the marks have
not been published. Also found were lithic resharpen-
ing flakes and a prismatic blade, part of a biface,
scra- pers, gravers, and cores (Breitburg et al. 1996).
Besides bones of mastodont, the site also contained
remains of deer, turtles, and muskrat.
Questions or issues? The bone marks thought to be
made by stone tools should be examined by experi-
enced taphonomic researchers who do not have a
stake in the current debate about the peopling of the
Americas, to provide a supporting voice to the argu-
ment that the animal bones and the lithics are behav-
iorally associated rather than unrelated members of a
palimpsest water-source assemblage.
Relation to Clovis? The lithics are not inconsistent
with forms known from Paleoindian assemblages,
but they do not show specific direct links to Classic
Clovis such as fluting or overface/ultrashot/overshot
flaking. The possible butchering of mastodont, along
with what may be either “background” or subsistence
faunal remains – deer, turtle, and muskrat – is mens, and more than four times as many
“cutmark- like” incisions on 29 specimens. Some
reminis- cent of Clovis mammoth sites in the Plains
marks were probably made by shovels and steel
and American West, such as Lehner, Murray trowels (see the
Springs,
8 mm film footage of the 1967 excavations,
Domebo, and Colby (Haynes 2002). As with recorded by L. Davis (ThisIsBozeman.com
Burning Tree, the site may reflect yet another [1967]; http://
manifes- tation of an occasional subsistence strategy
that even- tually culminated in a preference for the
largest mammals in some habitats (Haynes and
Hutson 2013; Surovell and Waguespack 2008, 2009;
Waguespack and Surovell 2003; but see Cannon and
Meltzer 2004, 2008; Grayson and Meltzer 2003 for dis-
senting opinions).

2.2.7 LINDSAY, MONTANA


Krasinski (2010) suggested that some mammoth
bones from the Lindsay site in Montana (Davis and
Wilson 1985; Hill and Davis 1998) had been cut,
chopped, and broken by people using stone tools. The
site yielded no flaked-stone lithics, but there were
several large sandstone blocks with the bones, which
Davis and Wilson (1985) had thought to be
manuports used as hammers or anvils to break open
bones. Earlier radiocarbon dating put the site in the
Clovis era (Hill and Davis 1998, for example);
however,
Waters and Stafford (2013) report new dates that
now indicate that it is about 1300–700 14C years
older than Clovis, depending on which assigned time
range for Clovis is preferred, or at 2 sigma roughly
1000–600 cal years older than the earliest Clovis ages
accepted here.
Questions or issues? Krasinski (2010) compared the
site’s mammoth bone marks to marks, which had
been experimentally produced by different processes.
To identify stone-tool-inflicted cuts, she applied
regression models that incorporated micromorpholo-
gical traits (e.g., mark length, width, and depth, apex
flatness, presence of concave or cracked walls and
apex striae, and type of termination of mark). Her
experiments and the statistical models showed that
stone-tool cuts have either V-shaped or flat apices
(that is, the bottoms of the marks), as do animal
tooth marks, and striae were present in the apices of
stone-tool cuts. Using regression models, the
accuracy of correct mark classifications (steel versus
stone versus tooth) varied, and was never 100 per
cent.
Thirteen marks on four Lindsay specimens had “a
>50 per cent probability to represent cutmarks pro-
duced by lithic implements” (Davis et al. 2013).
Also
noted, based on probabilities, were more than twice
as many steel-made marks on seven elements, more
than four times as many trample marks on 22 speci-
www.thisisbozeman.com/detail/digging-the-lindsey-
mammoth-1967/)). This diverse mix of marks made
by a wide variety of agents lends an air of ambiguity to
the site’s interpretation as a butchered mammoth.
Relation to Clovis? The dating, the uncertainty about
the bone marks, and the lack of lithics mean this find
can only be regarded as older than Clovis, but cannot be
unreservedly recorded as a memento of people who
were ancestral to Clovis. But, like Burning Tree, Coats-
Hines, and the Wisconsin mam- moths, if it is genuinely a
proboscidean butchery site, it shows an occasional human
preference for mega- fauna just before the Clovis era.

2.2.8 PAGE-LADSON, FLORIDA


The rich Page-Ladson site is under water in a 10-m deep
sinkhole in the Aucilla River, offering unusual
challenges – such as extremely low visibility in the deep
water – for the deciphering of its stratigraphy and
depositional processes (Dunbar 2006; Halligan
et al. 2013; Waters and Stafford 2013; Webb 2006).
“Abundant numbers of Paleoindian and early Archaic
artifacts from several time periods” have been recovered
from the sinkhole (Faught 2006,
173), including bifaces (Figure 7). During an early period
of use by animals and humans, before sea level rise at the
end of the Pleistocene, the site was a karst depression
situated almost 100 miles (ca. 162 km) from the Gulf
of Mexico, which is 5 miles (ca. 8 km) away today. A
deeply buried layer of sedi- ment contained a few lithic
flakes, a unifacial expedi- ent tool, and a possible
hammerstone, along with mastodont bones and wood
and seeds dated to
12,450 ± 40 14C yr BP (average of six dates) (ca.
14,600 cal yr BP at the 2-sigma range midpoint) (Dunbar
2006), which is about 1400–850 14C years older than the
Clovis time range in Waters and
Stafford (2007), or about 1200 cal years older than the
earliest Clovis date accepted here. A deposit of leaves,
bark, and wood is thought to be trampled mas- todont
dung, and contained a biface fragment (Halligan et al.
2013). A mastodont tusk from this level had six incised
grooves near the lipline, inter- preted as extraction marks
made by people removing
the tusk from the animal’s head (Webb 2006).
Questions or issues? Waters and Stafford (2013, 553)
think that the site’s data are “compelling evidence of early
human occupation,” but they also somewhat
contradictorily state that the early human presence is
not yet confirmed, because the stratigraphy is in need of
further study. The six deep grooves in the mastodont tusk
have not been as closely analyzed or adequately published
as they should be to support the interpretation as
extraction damage.
Relation to Clovis? The lithics from the deep level
reported in D u n b a r ( 2 0 0 6 ) do not show conspicuous
Figure 7 Bifaces from the Page-Ladson site (edited from an illustration in Stanford and Bradley 2012, figure 4.10; drawn by
Marcia Bakry; used by permission of D. Stanford).
CORRECTION: The biface on the far right is not a Page-Ladson
point.
and direct technological linkage to Classic Clovis. The object has not been extracted, but high-
The site’s possibly cut/chopped tusk is slim evidence resolution
for megamammal butchering, unless every possible X-ray computed tomography scanning shows that
non- the main length of the surviving shaft is
cultural source of the marks can be rigorously elimi- rectangular in
nated from consideration. Even so, the exploitation
of megafauna and creation of scattered sites at water
sources is suggestive of an emerging pre- or proto-
Clovis pattern in North America.

2.2.9 MANIS, WASHINGTON


At the Manis site in Washington, a mastodont rib
with a protruding osseous bit (Figure 8) made from a
mas-
todont bone or tusk (determined from DNA analysis)
was directly dated to 11,990 ± 30 14C yr BP (about
13,800 cal yr BP), almost 400 14C years older than
the Clovis horizon as accepted here. The pierced rib
was found in sediments left from a kettle pond, along
with fractured and flaked mastodont bones and puta-
tive cut marks on bone surfaces (Gilbow, unpub.
thesis 1981, cited in Gustafson et al. 1979; Waters
et al. 2011c; Runnings, unpub. thesis 1981, cited in
Waters et al. 2011c), which have not been confirmed
using more recently advanced analytical standards
(e.g., Dominguez-Rodrigo et al. 2010). No
unquestion- able flaked-stone artifacts were found
with the bones. A hydraulic bucket had exposed the
bones, splintering some elements in the process
(Figure 9).
Questions or issues? The protruding piece in the
Manis mastodont rib has a very small diameter.
Waters et al. (2011c), following Gustafson et al.
(1979), call it a projectile point, without any discus-
sion, although Stanford (1982, 209) had been more
equivocal, stating it was “tentatively identified as a
projectile point” and a “presumed projectile point.”
cross section rather than circular/rounded, with
maximum width and thickness estimated (from
published images with scale bars) of about 3 × 4 mm
at about 3.5 cm back from the conical tip. Judging
from the photographs and other images in Waters et
al. (2011c), the shaft of the Manis item has no traces of
grinding or smoothing, as is typical of bone piercing
implements from other places and times in prehistory.
Other undisputed bone points from arche-
ological contexts typically measure much larger in
shaft thicknesses. For example, “ivory points” from the
Yana site in Siberia (Nikolskiy and Pitulko 2013,
4195, figure 8) were rounded in cross section
( judging from the published images), with cross-sec-
tional diameters of about 6 mm, as measured 3.5 cm
from the tip; eyed needles from the site are smaller
(Pitulko et al. 2013, 32, figure 2.12I, J, L), but all
objects interpreted as weapons from the Yana site are
larger in diameter than the Manis specimen. Another
example of a projectile used against mammoth is
indicated at the Lugovskoe site in Siberia (Mashenko
et al. 2003; Zenin et al. 2003), where a mammoth
thoracic vertebra has a pierced hole measuring 7 × 10
mm across, roughly showing the original width of the
projectile point that entered the bone. This projectile
had a lithic inset, a fragment of which is still within the
hole. The projectile was hardly larger in cross-sectional
size than the Yana pieces without insets, and its
diameter was substan- tially larger than that of the
Manis piece.
Straight and curved ivory points or rods from
Alaska, Arizona, Florida, New Mexico, Ohio, Texas,
Washington, and Montana (Hemmings 2004, 185–189)
are all smoothed and polished, circular or oval to
semi-circular in cross section, and vary in
maximum diameters from 4.85 to 15 mm; the lengths
range up to 30 cm, and in one case 40 cm.
Figure 8 Two views of the Manis mastodont rib fragment with the embedded object (edited from a photographic montage in
Waters et al. 2011c, figure S4 in “supporting online material”; reprinted with permission from AAAS). Note the rectangular
cross section of the object, and also what may be additional bone material inserted into the rib or pushed out of it, alongside
the object.

A sample of complete or nearly complete ivory back from the tip, some of the so-called
points from Florida (see Hemmings 2004) are >7 to Bushman
>9 mm in diameter when measured 3.5 cm back
from the tips, much larger than the Manis piece’s
diameter and comparable to the Yana and
Lugovskoe points. An ivory sewing needle fragment
from Florida has a maximum cross-sectional diameter
of 4.8 mm at only 1.5 cm from the tapered tip, also
well above the diameter of the Manis piece
(Hemmings 2004, figure 3.16).
Similarities can be found between the Manis item
and the slender bone spear and arrow points from
southern Africa’s Later Stone Age and Iron Age
(Backwell et al. 2008, figures 3, 8). At about 3.5 cm
arrow points are 3–4 mm in diameter, although their
conical tips appear much sharper than the Manis
specimen, and they show signs of rounding and
shaping by human actions, unlike the Manis specimen.
These African specimens appear to be no longer than
about 13 cm, and were always bound into link shafts that
were inserted into reed shafts.
The Manis specimen’s small size and the evident
lack of human shaping to a more efficient circular cross
section should inspire a rigorous visual examin- ation of
the specimen and lead to experiments to test the
effectiveness of such a tiny projectile point.
Relation to Clovis? If the bone or ivory splinter is
indeed a projectile made by human hands, and the marked
and broken mastodont bones were part of a
Figure 9 The upper part of the figure shows Manis mastodont ribs that had been fragmented by mechanized excavation
equipment. Note the pebbles in colluvium above the bones and the stones in glacial till directly under the bones (photographed
by C. Vance Haynes in 1977, used with his permission). The lower part of the figure isthe bone-bed map (from Waters et al.
2011c, figure S3 in the “supporting online material,” edited to enlarge the text; note that the original figure had a misspelling of
one bone name,
“inominate;” reprinted with permission from AAAS.) with an added outline drawn around the approximate area in the photo
above.

butchered carcass, the site may be evidence that mega- indeed specifically targeted by human foragers in
fauna stressed by rapidly oscillating climates and com- some habitats (Haynes 2009).
peting for limited resources in the late glacial were
2.2.10 AYER POND, WASHINGTON a pond were directly dated 11,760 ± 70 and
At Ayer Pond, Washington, Bison antiquus bones 11,990 ± 25 14C yr BP (the latter date calibrated
found in clay and silt at the base of woody peat in at
2-sigma midpoint to about 13,850 cal yr BP) (Kenady understood – as a culture that made and used
et al. 2011). This is 100–390 14C years older than highly efficient tools
Clovis, or 450 cal years older than the earliest and whose foragers were not averse to leaving
Clovis date accepted here. No lithics were found, behind still-useful implements such as fluted
and no organic projectiles. Possible impact points on bifaces or dis- carding worn items such as
bones, unweathered-bone fracturing patterns, and unmodified flakes after using them.
chopmarks were also identified, but “no fine cut-
marks […] from slicing with a sharp edge” were
seen (Kenady et al. 2011, 135). No carnivore-gnaw
damage was found, and no evidence of animal-tram-
pling. Hence the materials are interpreted as a bison
dismembered and butchered by people, who dis-
carded low-utility body parts on the surface of a
frozen pond, and transported away the higher-utility
parts.
Questions or issues? The storage of large-mammal
body parts within frozen ponds is a recurring theme
in proboscidean-site interpretations by Fisher (e.g.,
see Fisher 1984a, 1984b, 1987, 1996, 2009), and is
equivocally suggested for the site by Kenady et al.
(2011), but a preferred twist is that the bison body
parts may have been cached on top of ice after butch-
ering, rather than under the ice in water. The bones in
illustrations appear to have been broken by impact
when in a fresh state, and were not carnivore
gnawed. Unfortunately, the discovery and collection
of the bones were not done by professionals, and the
discovery was made after a “tracked mechanical
exca- vator” (Kenady et al. 2011, 133) began digging
at the site; its weight could have deformed the
clay and
mucky silts enclosing the bones below peat
sediments, causing breakage and putting marks on
well-preserved elements. Workmen gathered up and
uncovered bones by hand after one bone was seen in
a cutwall of a mechanically opened excavation. The
bones were then stored for two years before a
professional was contacted and visited the find spot.
Kenady et al. (2011, 133) state that no shovels or
trowels were used by workmen to recover the
bones after mechanical
equipment exposed the first ones, based on workers’
memories two years later and apparently not on
formal notes or records.
Relation to Clovis? Interpretations of the site
support the idea that people in the centuries before
Clovis were competent foragers, who quickly and
effi- ciently killed and butchered large mammals, and
moved on without leaving much behind, somewhat
similar to Clovis behavior at some sites in the
American West. But the absence of associated lithic
and osseous implements makes it impossible to link
the find to Clovis as it has been classically
2.2.11 THE FIRELANDS GROUND SLOTH, OHIO excavators’ damage. However, Redmond et al. (2012)
At the Firelands Ground Sloth site, Ohio, a femur from an do report a variety of bone marks, which they
adult Jefferson’s ground sloth (Redmond et al. 2012) propose were probably made during handling and
was directly dated at 11,740 ± 35 14C yr BP (roughly storage, and also a type of mark called “claw
13,550 cal yr BP), 700 14C years older marks,” which are curved, broad, shallow, and U-
than the Waters and Stafford (2007) maximum age for shaped – and in my view very unlikely to have
Clovis, but only 140 14C years (150 cal years) older than been made on fresh mammalian cortical bone
the first-appearance date espoused in this paper. One surfaces by animal claws, which have not been shown
femur had dozens of fine incisions that were interpreted to perma- nently mark large-mammal cortical bone.
as made by unmodified and retouched stone flakes and The cut- marks themselves are said to have been
bifacial tools, used to chop or slice fresh bone surfaces. made by tools made of stone and not metal, based on
The sloth’s 10 skeletal elements had been given to features such as the deep and asymmetrical V-shaped
the Firelands Historical Museum in Norwalk, Ohio, but cross sec- tions with internal striations and steps, but
no detailed information is available except that they had these are also present in a proportion of metal-
been buried 4 ft deep in wet ground, probably a swamp or produced cuts (Krasinski 2010). Redmond et al.
peat bog. No lithics had been curated with the bones. (2012, 92) admit that the lack of information about
Upon learning about the bones, Redmond et al. (2012) the circumstances
initiated a taxonomic and taphonomic study of them. of recovery so long ago “preclude [sic] an absolute
Surface marks were inspected under optical instruments [italics in original] determination as to whether the
and a sample of silicon molds of the marks was examined incisions […] are the result of human butchering.”
in a scanning electron micro- scope by H. J. Greenfield. Relation to Clovis? Little can be said, other than
Small pieces of wood from inside bones were identified making the obvious (and hardly consequential)
as cf. conifer, consistent with expectable late Pleistocene acknowledgment that megafaunal butchering, pre-
flora. sumably following killing, is almost a commonplace
Questions or issues? The sloth bones had been dis- subsistence activity in the Clovis era and may have
covered decades before, so the methods of recovery, developed a few centuries before the appearance of
cleaning, and storage cannot be reviewed to eliminate the Classic Clovis lithic technology – if the sites discussed
possibility that the bone marks are preparators’ or immediately above have been correctly interpreted.

3. Discussion dently and directly dated, but if there is a


14 confirmable
One-thousand C years may approach the maximum
temporal association with the dated (older than
spread of time in which techno- and socio-cultural
traits among low-density foragers can be expected to Clovis) mastodont dung deposit, they very well
could be closely ancestral to Clovis. The rest of the
show clear remaining hints of an ancestral–
sites reviewed here either have very
descendant relationship. The sites reviewed here
lengthyseparations of dates on their pre-Clovis and
reasonably
Clovis lithics (Topper), no Clovis at all overlying
suggest a human presence in many parts of the
the pre-Clovis levels (Meadowcroft, Paisley Caves),
continent well before the Clovis era, even though
nondescript or undiagnostic lithics in earliest levels
they are not without interpretive problems. But only a
(Big Eddy, Hebior, Schaefer, and Coats-Hines), or
very few have what can be considered even indirect
no lithics at all (Lindsay, Manis, and Ayer Pond),
technological links to Classic Clovis, or to any other
making inferences about possible lineal relationships
defined cultural
to Clovis unwarranted. And even with the
“type” such as Western Stemmed. Leading the list of
occasional lithic technological similarities between
sites that could be directly ancestral to Clovis is the
Classic Clovis and some pre-Clovis assemblages, it
Debra L. Friedkin site in Texas, where the lithic
is difficult to make rigorous and unambiguous
assem- blage is very similar to what has been found
reconstructions of antecedent or coexisting
at the Aubrey Clovis site, also in Texas (Ferring
relationships, because special- ized lithic
2001, 2012; Morrow et al. 2012), although Jennings
technological theories are not on hand to guide such
and Waters (2014) point out differences, too. Other
reasoning, as Waguespack (2007) and Shott (2013)
possible lineal relationships can be seen in parts of
have pointed out. Anderson et al. (2013, 185)
the eastern Beringian sites (Goebel et al. 1991), but
with major resignedly recognized that “at present all we know
is
differences there, too, and possibly in Meadowcroft’s
that appreciable variability characterizes pre-Clovis,
early assemblages and the pre-Clovis material from
Clovis, and immediate post-Clovis industries in
Cactus Hill, but the technological relationships are
North America, even among chipped-stone
tenuous at best. Page-Ladson’s bifaces are not all
bifaces.”
confi-
Contributing to the confusion is the unfortunate
scarcity the archeological literature. Some of these models see
of detailed publications about many of the sites, Clovis as the sign of a separate human dispersal,
which would provide essential information, the rather latecomers to the Americas, while others see
critical data points, working hypotheses, defended Clovis as only a set of technological ideas that
interpretations, responses to circulating critiques, and spread among long-established human subpopu-
so forth. lations. Some of the disagreement may arise from
different degrees of documentation that can be found
3.1 Disparate models about proposed sites. In too many cases, there are
Several different models of how Clovis developed from not fully developed integrated descriptions of
pre-existing populations south of Beringia are found material culture (if any has been found), dating, and
in geology, and reasonable counter-arguments are
sometimes not well made when the claims are
doubted that sites were humanly made.
The “Staging Area” model (Anderson 1996) was a
proposal that a founding population in the lower 48
states, originators of Clovis culture, fragmented into
separate daughter subpopulations, which adapted to
different micro- and macro-environments in North
America, although these subpopulations continued
to share some “Clovis core” features (Anderson and
Gillam 2000; Smallwood 2012). Fluting and other
Clovis-specific features such as efficient biface thin-
ning by ultrashot or overface flaking and macroblade
manufacture may have been developed before the
staging-area phase, originating in a founding popu-
lation before it fragmented. This is essentially a
“Clovis-first” model. The staging areas continued to
be used even after Clovis was replaced by early
Archaic complexes.
A different kind of model was outlined by Waters
and Sta fford (2007), who contended that the
surpris- ingly rapid spread of fluting and other
features was the result of adoption of the technology
by pre-existing and long-resident human
subpopulations. The ulti- mate place of invention of
the technology is unknown. Collins et al. (2013)
asserted a long sequence of human occupation in
the Americas, starting almost 30,000 14C yr BP and
exhibiting seven divergent pat- terns of lithic
production technology; the message is that different
cultures originating at different times and places
were in the Americas well before the LGM. Note
that dates for some of the recognized pat- terns are
widely separated, such as in Pattern 1 with
bracketing dates >27,000 and <12,000 14C yr BP,
and Pattern 5 with bracketing dates >20,000 and
<11,000 14C yr BP, but without sites showing dating
gradients between the extremes. This may suggest tech-
nological convergence by unrelated cultures and not
lineal connection of the technologies.
MacNeish (2004) also described multiple techno-
logical patterns in a much longer time span for the
Americas, beginning around 100 ka with a Lower
Paleolithic kind of Chopper-Chopping Tool stage of
human dispersal in the Americas, progressing
through a Zhoukoudian Upper Cave-like Unifacial
stage of human presence, and ending with a stage of
leaf-shaped points, burins, and blades, which includes continent. It also does not take into consideration
microblades. These were separate waves of important points about late Pleistocene human
movement (MacNeish 1971, 2004) that represent foraging, such as oscillating climate changes
evolutionary stages in hominid technology. that were affecting
Adovasio and Pedler (2013, 513) asserted that at
least 30 archeological sites in the Americas have sup-
portable claims for being older than Clovis, and con-
cluded that “the diversity of tool kits and the
attendant lifestyles […] [at the older-than-Clovis
sites] point toward multiple movements by different
groups at different times via different routes into the
New World well before the Clovis temporal bench-
mark” (Adovasio and Pedler 2013, 518). They did
not try to reconstruct the origins of Clovis out of all
this diversity.
There are yet other possibilities about Clovis
origins, which could be summarized here, if more
space was available – such as the implications in
Beck and Jones (2010) that the unconnected
cultures
making Western Stemmed and Clovis points des-
cended from different ancestors who entered the
American West from different directions, both being
almost the same age. In this proposal, the stemmed
points originated among far western (coastal)
migrants, while the fluted points came from a popu-
lation originating far inland, to the southeast. The
sep- arate ancestry of the fluted-point population is
not suggested. However, genetic evidence from
modern and ancient DNA indicates a single Native
American origin in Northeast Asia, and a single
migration and expansion south of the ice sheets.
My own view is that rapid human spread carrying
the Clovis technology through an empty or nearly
empty continent cannot be ruled out as a possibility,
and it becomes more tenable if a duration of 1000
years was involved in the dispersal. The successful
dis- persal of a very small founding population into
wood- lands, steppes, and grasslands south of the ice
sheets may have been accomplished in just centuries,
as human demes found the game trails, faunal
refugia, and competition-free open ranges ahead.
Meltzer (2009, 231) has stated that “most studies
agree detailed landscape learning is a prolonged
process, with estimates ranging from 200–1000
[calen- dar] years.” An implication is that centuries
of human
invisibility or near invisibility may have passed
before the earliest people in any region accumulated
enough material debris anywhere to create what
would be dis- coverable as sites thousands of years
later. This makes sense, but it does not explain how
the new Clovis technology could have spread so
quickly through the near-invisible different cultural
groups who were presumably spread thinly across the
large-mammal populations, undoubtedly making them were unconvinced about Monte Verde, Chile – prob-
easier to track and attack, even if lowering their general ably the best-described and most publicized pre-
continental densities, and the absence of human Clovis site, familiar to every professional
competition for resources such as high-quality toolstone archeologist. Claims made for lesser known sites may
and water. The first American foragers be even less convincing to professional archeologists
may not have needed 200–1000 years to frame who have not been directly involved with them. Such
mental maps of the open ranges around them, unlike the questioning bedevils the consensual acceptance of
early English colonists mentioned by Meltzer (2009, 231) possible patterns in human dispersals and brings the
who tried to make a point about the slow- ness of relevance of some sites into question when it comes
landscape learning: in spite of being techno- logically to reviews such as
advanced, the English were ill-equipped as farmers, faced this one – are the claims worth taking seriously if
human competition for resources, and suffered through they do not convince a substantial number of pro-
monumental drought events, conse- quently requiring fessionals? Should they be incorporated into models
more than a century to produce accurate paper maps. And of the first human entry into the Americas, or are
it is also worth noting that the English newcomers they better left out of the discussion?
essentially stopped at the eastern American Fall Line for a I think some interpretive disagreements about
century or so, while French explorers far more quickly older- than-Clovis claims probably could be
traveled much deeper into the unfamiliar continent. minimized, if scrupulously complete and integrated
publications were to provide details about
4. Conclusion stratigraphy, dating, geology, materials recovered,
and hypotheses of poten- tial meanings. Many sites in
Let us […] work and wedge our feet downward
the samples above do not have such well-thought-out
through the mud and slush of opinion, and preju-
publications, and the issues raised about the claims
dice, and tradition, and delusion, and appear- ance
[…] till we come to a hard bottom […] which we are still unresolved in the minds of scholars. It is also
can call reality […] [H. D. Thoreau, 1854: Walden, imperative that when- ever skeptical colleagues
or Life in the Woods] question the claims and interpretations, they are
addressed methodically and objectively. If criticisms
In the Wheat (2012) survey of professional opinions are chronically suppressed or
about early American sites, 33 per cent of respondents
dismissed, combative partisans of opposing interpre- were too scattered and transitory in any one locality
tations begin an unproductive era of verbal head-butting. to leave much of an archeological record, as Meltzer
The future discourse can perhaps be moderated in (2009) and others have surmised. But by the time of
those ways, but what about the material ambiguity Clovis, the record had become very visible and dis-
that may remain about certain finds, the type of incon- tinct. The ancestral humans of 1000 years before
clusiveness never to be resolved perfectly? For Clovis were in the process of establishing an
example, stone-tool cutmarks sometimes have the archeolo- gical footprint, even if they left hardly any
same features as marks created by steel-tool edges, as discover- able traces for centuries. Our picture of
mentioned above in the discussion about the Lindsay the transition from the older-than-Clovis period to
mammoth site. If analysis of a site’s marked-up the Clovis era may always remain indefinite and
bones shows that a small percentage of all examined perplexing because of the visibility problem. In the
marks statistically appear to be likely stone-tool cuts, assessment of some scholars, if there had been
and the rest are of unknown or noncultural origins, human populations in the New World many more
should the evidence for cutting by stone tools be than 1000 years before Clovis, the archeological
given greater weight than the possibility that a small traces of the millennium before Clovis, as reviewed
fraction of noncultural marks simply possess the fea- above, should be much clearer and more extensive
tures of stone-tool cuts? than it is, which seems to weaken the hypothetical
There is also potential ambiguity about spatial case for a very early human occupation of the New
associations of materials in some sites’ sedimentary World.
environments, such as vertisols or unconsolidated Over a decade ago, Yesner et al. (2004, 212–213)
sands, and expert opinions expectably differ about asked “What do we know about Paleoindian
downward creep or post-depositional displacement. coloniza- tions,” but no new facts have been
In cases such as these, new data or new evidence discovered since then. The only facts then and now
from the sites that could resolve the disagreements are just these two:
may never be coming, and the opinions will go on first human beings in the Americas originated from
clashing for a long time. Asian populations, and if all the claims are believed,
Probably the very first humans in the New World there was a great deal of variability in regional
subsis- tence and technologies. Anything else which
and ambiguous” (Shott 2013, 150), and incompletely
can be confidently concluded about Clovis origins is
preserved. Exasperating questions remain, which will
limited, because the archeological record has not
continue to inspire the testing of new and evolving
been fully stories about the peopling of the Americas.
decoded and the evidence itself is “sparse, scattered,

5. Acknowledgements
Thanks to Ted Goebel for inviting this review.
Thanks also to the late Paul Martin for showing how
to deal with interpretive disagreements, even the most
aggres- sive kind, and who thoughtfully tried to teach
so many students about late-glacial North America. I
wish his patience and civility had been learned by
more people in this profession. I also want to thank
William Gardner who infected me with the Clovis
virus, and C. Vance Haynes (no biological relation
that can be traced) for sharing time and knowledge
over so many years. Special thanks go to Vance
Holliday, Stuart Fiedel, David Kilby, Vance Haynes,
and a “somewhat anonymous reviewer,” in the
journal editor’s words, for reading a draft of this
paper and making major and minor suggestions to
improve it, and to Janis Klimowicz for essential help
in its preparation.

References
Aceituno, F. J., N. Loaiza, M. E. Delgado-Burbano, and
G. Barrientos. 2013. “The initial human settlement of north-
west South America during the Pleistocene/Holocene tran-
sition: Synthesis and perspectives.” Quaternary International
301: 23–33.
Adovasio, J. M. 1993. “The ones that will not go away: A biased
view of the pre-Clovis populations in the New World.” In
From Kostenki to Clovis, edited by O. Soffer, and
N. D. Praslov, 199–218. New York: Plenum Press.
Adovasio, J. M., J. Gunn, J. Donahue, and R. Stuckenrath.
1978. “Meadowcroft Rockshelter, 1977: An overview.”
American Antiquity 43: 632–651.
Adovasio, J. M., J. D. Gunn, J. Donahue, R. Stuckenrath, with sec-
tions by Applegarth, J. D., R. C. Carlisle, D. T. Clark,
J. Donahue, D. Faingnaert, J. Guilday, W. C. Johnson,
D. Krinsley, K. Lord, E. Skirboli, P. G. Wiegman. 1977.
“Meadowcroft Rockshelter: Retrospect 1976.” Pennsylvania
Archaeologist 47(2–3): 1–93.
Adovasio, J. M., and J. Page. 2002. The First Americans: In Pursuit
of Archaeology’s Greatest Mystery. New York: Random House.
Adovasio, J. M., and D. Pedler. 2005. “A long view of deep time at
Meadowcroft Rockshelter.” In Paleoamerican Origins: Beyond
Clovis, edited by R. Bonnichsen, B. T. Lepper, D. Stanford,
and M. Waters, 23–28. College Station: Center for the Study
of the First Americans, Texas A&M University.
Adovasio, J. M., D. Pedler, J. Donahue, and R. Stuckenrath. 1999.
“No vestige of a beginning nor prospect of an end: Two
decades of debate on Meadowcroft Rockshelter.” In Ice Age
People of North America: Environments, Origins, and
Adaptations, edited by R. Bonnichsen and K. L. Turnmire,
416–431. Corvallis: Center for the Study of the First
Americans,
Oregon State University Press.
Adovasio, J. M., and D. R. Pedler. 2013. “The ones that still won’t
go away: More biased thoughts on the pre-Clovis peopling of
the New World.” In Paleoamerican Odyssey, edited by
K. E. Graf, C. V. Ketron, and M. R. Waters, 511–520.
College Station: Center for the Study of the First Americans,
Texas A&M University.
Ahler, S. A. 2000. “Use-wear evaluation of five rock specimens.”
In The 1999 Excavations at the Big Eddy Site (23CE426),
edited by N. H. Lopinot, J. H. Ray, and M. D. Conner, 230–
236. Southwest Missouri State University Center for
Archaeological Research. Special Publication No. 3.
Amick, D. S. 2004. “A possible ritual cache of Great Basin R. S. MacNeish and J. G. Libby, 417–430. Albuquerque:
stemmed bifaces from the terminal Pleistocene–early University of New Mexico Press.
Holocene occu- pation of NW Nevada, USA.” Lithic
Technology 29(2): 119–145.
Anderson, A. D. 1962. “The Cooperton mammoth: A preliminary
report.” Plains Anthropologist 7: 110–112.
A. D. Anderson (ed.). 1975. “The Cooperton mammoth: An early
man bone quarry.” Great Plains Journal 14: 130–173.
Anderson, D. G. 1996. “Models of Paleoindian and Early Archaic
settlement in the lower Southeast.” In The Paleoindian and
Early Archaic Southeast, edited by D. G. Anderson and K.
E. Sassaman, 29–57. Tuscaloosa: University of Alabama
Press. Anderson, D. G., T. G. Bissett, and S. J. Yerka. 2013.
“The late-
Pleistocene human settlement of interior North America: The
role of physiography and sea-level change.” In
Paleoamerican Odyssey, edited by K. E. Graf, C.
V. Ketron, and
M. R. Waters, 183–203. College Station: Center for the Study
of the First Americans, Texas A&M Press.
Anderson, D. G., and J. C. Gillam. 2000. “Paleoindian colonization
of the Americas: Implications from an examination of physio-
graphy, demography, and artifact distribution.” American
Antiquity 65: 43–66.
Anovitz, L. M., J. M. Elam, L. R. Riciputi, and D. R. Cole. 1999.
“The failure of obsidian hydration dating: Sources, impli-
cations, and new directions.” Journal of Archaeological
Science 26: 735–752.
Backwell, L., F. d’Errico, and L. Wadley. 2008. “Middle Stone
Age bone tools from the Howiesons Poort layers, Sibudu
Cave, South Africa.” Journal of Archaeological Science
35: 1566–1580.
Beck, C., and G. T. Jones. 1997. “The terminal Pleistocene/early
Holocene archaeology of the Great Basin.” Journal of World
Prehistory 11: 161–236.
Beck, C., and G. T. Jones. 2010. “Clovis and Western Stemmed:
Population migration and the meeting of two technologies in
the Intermountain West.” American Antiquity 75: 81–116.
Boëda, E., A. Lourdeau, C. Lahaye, G. Daltrini Felice, S. Viana,
I. Clemente-Conte, M. Pino, M. Fontugne, S. Hoeltz,
N. Guidon, A.-M. Pessis, A. Costa Da, and M. Pagli. 2013.
“The late-Pleistocene industries of Piauí, Brazil: New data.”
In Paleoamerican Odyssey, edited by K. E. Graf,
C. V. Ketron, and M. R. Waters, 445–465. College Station:
Center for the Study of the First Americans, Texas A&M
Press.
Bradley, B. A. 1993. “Paleo-Indian flaked stone technology in the
North American High Plains.” In From Kostenki to Clovis:
Upper Paleolithic-Paleo-Indian Adaptation, edited by
O. Soffer and N. D. Praslov, 251–262. New York: Plenum
Press. Bradley, B. A., M. B. Collins, and C. A. Hemmings. 2010.
Clovis Technology. Ann Arbor: International Monographs in
Prehistory.
Breitburg, E., J. B. Broster, A. L. Reesman, and R. G. Stearns.
1996. “The Coats-Hines site: Tennessee’s first Paleoindian-
mastodon association.” Current Research in the Pleistocene
13: 6–8.
Bryan, A. L. 1988. “The relationship of the stemmed point and
fluted point traditions in the Great Basin.” In Early Human
Occupation in Far Western North America: The Clovis–
Archaic Interface, edited by J. A. Willig, C. M. Aikens, and J.
L. Fagan, 53–74. Nevada State Museum. Anthropological
Papers Number 21.
Bryan, A. L., and D. Tuohy. 1999. “Prehistory of the Great Basin/
Snake River Plain to about 8,500 years ago.” In Ice Age
Peoples of North America: Environments, Origins, and
Adaptations of
the First Americans, edited by R. Bonnichsen and
K. L. Turnmire, 249–263. Corvallis: Oregon State University
Press and Center for the Study of the First Americans.
Cannon, M. D., and D. J. Meltzer. 2004. “Paleoindian foraging:
Examining the faunal evidence for large mammal
specialization and regional variability in prey choice.”
Quaternary Science Reviews 23: 1955–1987.
Cannon, M. D., and D. J. Meltzer. 2008. “Explaining variability in
early Paleoindian foraging.” Quaternary International 191: 5–
17.
Chrisman, D., R. S. MacNeish, J. Mavalwala, and H. Savage.
1996. “Late Pleistocene human friction skin prints from
Pendejo Cave, New Mexico.” American Antiquity 61(2): 357–
376.
Chrisman, D., J. Mavalwala, H. Savage, and A. Tessarolo. 2003.
“Friction skin imprints and hair.” In Pendejo Cave, edited by
Clague, J. J., T. A. Ager, and R. W. Mathewes. 2004. “Environments of 2: The Archaeological Context and Interpretation. Washington,
northwestern North America before the Last Glacial Maximum.” In DC: Smithsonian Institution Press.
Entering America: Northeast Asia and Beringia before the Dillehay, T. 1999. “The late Pleistocene cultures of South America.”
Last Glacial Maximum, edited by Evolutionary Anthropology 7: 206–216.
D. B. Madsen, 63–94. Salt Lake City: University of Utah Press. Dillehay, T. D. 2000. “Preliminary micro-use-wear of four pre-
Collins, M., and J. Lohse. 2004. “The nature of Clovis blades and Clovis-age objects.” In The 1999 Excavations at the Big Eddy
blade cores.” In Entering America: Northeast Asia and Beringia Site (23CE426), edited by N. H. Lopinot, J. H. Ray, and
before the Last Glacial Maximum, edited by M. D. Conner, 221–229. Southwest Missouri State University
D. B. Madsen, 159–183. Salt Lake City: University of Utah Press. Center for Archaeological Research. Special Publication No. 3.
Collins, M. B. 2007. “Discerning Clovis subsistence from stone arti- Dillehay, T. D. 2013. “Entangled knowledge: Old trends and new
facts and site distributions on the southern Plains periphery.” In thoughts in first South American studies.” In Paleoamerican
Foragers of the Terminal Pleistocene in North America, edited by Odyssey, edited by K. E. Graf, C. V. Ketron, and
R. B. Walker and B. N. Driskell, 59–87. Lincoln: University of M. R. Waters, 377–395. College Station: Center for the Study
Nebraska Press. of the First Americans, Texas A&M University.
Collins, M. B., D. J. Stanford, D. L. Lowery, and B. A. Bradley. 2013. Dillehay, T. D., M. Pino, J. Rossen, C. Ocamp, P. Rivas, D.
“North America before Clovis: Variance in temporal/ spatial Pollack, and G. Henderson. 1999. “Reply to Fiedel, Part I.”
cultural patterns, 27,000–13,000 cal yr BP.” In Paleoamerican Scientific American Discovering Archaeology Special Report:
Odyssey, edited by K. E. Graf, C. V. Ketron, and M. R. Waters, Monte Verde Revisited: 12–14.
521–539. College Station: Center for the Study of the First Dillehay, T. D., C. Ramirez, M. Pino, M. B. Collins, J. Rossen, and
Americans, Texas A&M University. J. D. Pino-Navarro. 2008. “Monte Verde: Seaweed, food,
Davis, L. B., C. L. Hill, and K. Krasinski. 2013. “Evidence for pre- medi- cine, and the peopling of South America.” Science 320:
Clovis human activity associated with a mammoth in late 784–786.
Pleistocene eastern Montana.” Poster presented at Paleoamerican Dixon, E. J. 2013. Arrows and Atl Atls: A Guide to the Archeology of
Odyssey: A Conference Focused on First Americans Archaeology, Beringia. Anchorage: US Department of the Interior, National
Santa Fe, New Mexico, 16–19 October. Park Service.
Davis, L. D., and M. C. Wilson. 1985. “The late Pleistocene Lindsay Dominguez-Rodrigo, M., T. R. Pickering, and H. T. Bunn. 2010.
Mammoth (24DW501), eastern Montana: Possible man–mam- moth “Configurational approach to identifying the earliest hominin
association.” Current Research in the Pleistocene 2: 97–98. Deter-Wolf, butchers.” Proceedings of the National Academy of Sciences
A., J. W. Tune, and J. B. Broster. 2011. “Excavations and dating of late (USA) 107: 20929–20934.
Pleistocene and Paleoindian deposits at the Coats-Hines site, Dominguez-Rodrigo, M., T. R. Pickering, and H. T. Bunn. 2012.
Williamson County, Tennessee.” Tennessee “Experimental study of cut marks made with rocks unmodified
Archaeology 5(2): 142–156. by human flaking and its bearing on claims of ∼3.4-million-
Dickinson, W. R. 2011. “Geological perspectives on the Monte year-old butchery evidence from Dikika, Ethiopia.” Journal
Verde archaeological site in Chile and pre-Clovis coastal migration in of Archaeological Science 39: 205–214.
the Americas.” Quaternary Research 76: 201–210. Dillehay, T. 1989. Driese, S. G., L. C. Nordt, M. R. Waters, and J. L. Keene. 2013.
Monte Verde: A Late Pleistocene Site in Chile. “Analysis of site formation history and potential disturbance
Volume 1. Paleoenvironment and Site Context. Washington, DC: of stratigraphic context in vertisols at the Debra L. Friedkin
Smithsonian Institution Press. archaeological site in central Texas, USA.” Geoarchaeology
Dillehay, T. 1997. Monte Verde: A Late Pleistocene Site in Chile. Volume 28: 221–248.

Duke, D. 2011. If the Desert Blooms: A Technological Perspective on http://www.archaeologysouthwest.org/what-we-


Paleoindian Ecology in the Great Basin from the Old River Bed, do/information/ exhibits/peo/pre-clovis/).
Utah. PhD dissertation, Department of Anthropology, Feathers, J. K., E. J. Rhodes, S. Huot, and J. M. McAvoy. 2006.
University of Nevada, Reno. “Luminescence dating of sand deposits related to late
Dunbar, J. S. 2006. “Paleoindian archaeology.” In First Floridians Pleistocene human occupation at the Cactus Hill site,
and Last Mastodonts: The Page-Ladsen Site in the Aucilla Virginia, USA.” Quaternary Geochronology 1: 167–187.
River, edited by S. D. Webb, 403–435. Dordrecht: Springer. Ferring, C. R. 2001. Archaeology and Paleoecology of the
Ellis, C. 2004. “Hi-Lo: An early lithic complex in the Great Lakes Aubrey
region.” In The Late Paleo-Indian Great Lakes: Geological Clovis Site (41DN479), Denton County, Texas. Denton:
and Archaeological Investigations of Late Pleistocene Center for Environmental Archaeology, University of North
and Texas.
Early Holocene Environments, edited by L. J. Jackson and A. Ferring, R. 2012. “The ‘Long’ Clovis Chronology: Evidence
A. Hinshelwood, 57–83. Canadian Museum of Civilization from the Aubrey and Friedkin Sites, Texas.” Paper presented
Mercury Series Archaeology Paper 165. at the 77th Annual Meeting of the Society for American
Eren, M. I., R. J. Patten, M. J. O’Brien, and D. J. Meltzer. 2013. Archaeology, Memphis, Tennessee, April 18–22.
“Refuting the technological cornerstone of the Ice-Age Fiedel, S. J. 1999. “Artifact provenience at Monte Verde:
Atlantic crossing hypothesis.” Journal of Archaeological Confusion and contradictions.” Scientific American
Science 40: 2934–2941. Discovering Archaeology Special Report: Monte Verde
Eren, M. I., R. J. Patten, M. J. O’Brien, and D. J. Meltzer. 2014. Revisited, 1–12.
“More on the rumor of ‘intentional overshot flaking’ and the Fiedel, S. J. 2014. “Did pre-Clovis people inhabit the Paisley
purported Ice-Age Atlantic crossing.” Lithic Technology Caves (and why does it matter)?” Human Biology 86(1),
39(1): 55–63. Article 8.
Fiedel, S. J. n.d. “Fiedel’s response to ‘Reply to Fiedel, Part I’.”
Erlandson, J. M. 2013. “After Clovis-First collapsed: Reimagining (https://www.academia.edu/3304654/Fiedels_Response_to_
the peopling of the Americas.” In Paleoamerican Odyssey,
Reply_to_Fiedel_Part_1_Dillehay_et_al._2000, accessed 6
edited by K. E. Graf, C. V. Ketron, and M. R. Waters,
July 2014).
127–132. College Station: Center for the Study of the First
Americans and Texas A&M. Fiedel, S. J., and J. E. Morrow. 2012. “Comment on ‘Clovis and
Western Stemmed: Population migration and the meeting of
Fariña, R. A., S. Tambusso, L. Varela, A. Czerwonogora, M. Di two technologies in the Intermountain West’ by Charlotte
Giacomo, M. Musso, R. Bracco, and A. Gascue. 2014.
Beck and George T. Jones.” American Antiquity 77(2): 376–
“Arroyo del Vizcaíno, Uruguay: A fossil-rich 30-ka-old mega- 385.
faunal locality with cut-marked bones.” Proceedings of the Fisher, D. C. 1984a. “Taphonomic analysis of late Pleistocene
Royal Society B 281: 20132211. mas-
Faught, M. K. 2006. “Paleoindian archaeology in Florida and todont occurrences: Evidence of butchery by North
Panama: Two circumgulf regions exhibiting waisted American Paleo-Indians.” Paleobiology 10: 338–357.
lanceolate projectile points.” In Paleoindian Archaeology: A
Fisher, D. C. 1984b. “Mastodon butchery by North American
Hemispheric Perspective, edited by J. E. Morrow and C. Paleo-Indians.” Nature 308: 271–272.
Gnecco, 164–183. Gainesville: University Press of Florida.
Fisher, D. C. 1987. “Mastodont procurement by Paleoindians of
Faught, M. K., and A. K. L. Freeman. 1998. “Paleoindian com- the Great Lakes region: Hunting or scavenging?” In The
plexes of the terminal Wisconsin and early Holocene.” In Evolution of Human Hunting, edited by M. H. Nitecki, and
Paleoindian and Archaic Sites in Arizona, edited by D. V. Nitecki,
J. B. Mabry, 33–52. Phoenix: Arizona State Historic 309–421. New York: Plenum.
Preservation Office (accessed online 19 November 2014 at Fisher, D. C. 1996. “Extinction of proboscideans in North
America.” In The Proboscidea: Evolution and Palaeoecology
of Elephants and Their Relatives, edited by J. Shoshani and
P. Tassy, 296–315. Oxford: Oxford University Press.
Fisher, D. C. 2009. “Paleobiology and extinction of proboscideans
in the Great Lakes region of North America.” In American
Megafaunal Extinctions at the End of the Pleistocene, edited
by G. Haynes, 55–75. Dordrecht: Springer.
Fisher, D. C., B. T. Lepper, and P. E. Hooge. 1991. “Taphonomic
analysis of the Burning Tree mastodon.” Current Research in
the Pleistocene 8: 88–92.
Fisher, D. C., B. T. Lepper, and P. E. Hooge. 1994. “Evidence for
butchery of the Burning Tree Mastodon.” In The First
Discovery of America: Archaeological Evidence of the
Early
Inhabitants of the Ohio Area, edited by W. S. Dancey, 43–57.
Columbus: The Ohio Archaeological Council.
Fitzgibbon, P. T., “with the assistance of Herbstritt, J.,
W. C. Johnson, C. Robbins”. 1982. “Lithic artifacts from
Meadowcroft Rockshelter and the Cross Creek drainage.” In
Meadowcroft: Collected Papers on the Archaeology
of
Meadowcroft Rockshelter and the Cross Creek Drainage,
edited by R. C. Carlisle and J. M. Adovasio, 91–111.
Prepared for the symposium “The Meadowcroft Rockshelter
Rolling Thunder Review: Last Act,” 47th Annual Meeting of
the Society for American Archaeology, Minneapolis, April
14–17.
Flannery, T. 2003. “Who came first?” The New York Review of
Books 50(1): 51–53.
Frison, G. C., ed. 1996. The Mill Iron Site. Albuquerque: University
of New Mexico Press.
Gilbert, M. T. P., D. L. Jenkins, A. Götherstrom, N. Naveran, J.
J. Sanchez, M. Hofreiter, P. F. Thompson, J. Binladen,
T. F. G. Higham, R. M. Yohe, II, R. Parr, L. S. Cummings,
and E. Willerslev. 2008. “DNA from pre-Clovis human copro-
lites in Oregon, North America.” Science 320: 786–789.
Goebel, T. 2004. “The search for a Clovis progenitor in subarctic
Siberia.” In Entering America: Northeast Asia and Beringia
before the Last Glacial Maximum, edited by D. B. Madsen,
311–356. Salt Lake City: University of Utah Press.
Goebel, T., and I. Buvit. 2011. “Introducing the archaeological
record of Beringia.” In From the Yenisei to the Yukon:
Interpreting Lithic Assemblage Variability in Late
Pleistocene/
Early Holocene Beringia, edited by T. Goebel and I. Buvit, 1–
30. College Station: Texas A&M University Press.
Goebel, T., and J. L. Keene. 2014. “Are Great Basin stemmed points
as old as Clovis in the Intermountain West? A review of the geo-
chronological evidence.” In Archaeology for All Times:
Papers in Honor of Don D. Fowler, edited by J.
Janetski and
N. Parezo, 35–60. Salt Lake City: University of Utah Press.
Goebel, T., R. Powers, and N. Bigelow. 1991. “The Nenana
Complex of Alaska and Clovis origins.” In Clovis Origins and
Adaptations, edited by R. Bonnichsen and K. L. Turnmire,
49–79. Corvallis: Center for the Study of the First Americans,
Oregon State University.
Goebel, T., H. L. Smith, L. DiPietro, M. R. Waters, B. Hockett,
K. E. Graf, R. Gal, S. B. Slobodin, R. J. Speakman, S.
G. Driese, and D. Rhode. 2013. “Serpentine Hot Springs,
Alaska: Results of excavations and implications for the age
and significance of northern fluted points.” Journal of
Archaeological Science 40: 4222–4233.
Goebel, T., M. R. Waters, and D. H. O’Rourke. 2008. “The late
Pleistocene dispersal of modern humans in the Americas.”
Science 319: 1497–1502.
Goldberg, P., F. Berna, and R. Macphail. 2009. “Comment on
‘DNA from pre-Clovis human coprolites in Oregon, North
America’.” Science 325: 148-c.
Goodyear, A. C. 2005. “Evidence of pre-Clovis sites in the eastern
United States.” In Paleoamerican Origins: Beyond Clovis,
edited by R. Bonnichsen, B. Lepper, D. Stanford, and
M. Waters, 103–112. College Station: Texas A&M University
Press.
Graf, K. E. 2008. Uncharted Territory: Late Pleistocene Hunter-
Gatherer Dispersals in the Siberian Mammoth-Steppe. PhD
dis- sertation, Department of Anthropology, University of
Nevada, Reno.
Graf, K. E. 2009. “‘The Good, the Bad, and the Ugly’: Evaluating
the radiocarbon chronology of the middle and late Upper
Paleolithic of the Enisei River Valley, south-central Siberia.”
Journal of Archaeological Science 36(3): 694–707.
Grayson, D. K., and D. J. Meltzer. 2003. “A requiem for North
American overkill.” Journal of Archaeological Science 30: 585–593.

Gustafson, C. E., D. W. Gilbow, and R. D. Daugherty. 1979. “The Haynes, G., and K. E. Krasinski. 2010. “Taphonomic
Manis mastodon site: Early man on the Olympic Peninsula.” fieldwork in southern Africa and its application in
Canadian Journal of Archaeology/Jounal Canadien studies of the earliest
d’Archéologie 3: 157–164.
Halligan, J., J. S. Dunbar, B. Fenerty, and E. Green. 2013.
“Submerged Paleoindian Sites in the Aucilla River of
Northwestern Florida: New Geoarchaeological and
Archaeological Research.” Poster presented at the
Paleoamerican Odyssey Conference, Santa Fe, New Mexico,
October 16–19.
Hand, D. J. 2014. The Improbability Principle: Why Coincidences,
Miracles, and Rare Events Happen Every Day. New York:
Scientific American/Farrar, Straus and Giroux.
Haynes, C. V., Jr. 1964. “Fluted projectile points: Their age and
dis- persion.” Science 145: 1408–1413.
Haynes, C. V., Jr. 1966. “Elephant hunting in North America.”
Scientific American 214: 104–112.
Haynes, C. V., Jr. 1967. “Carbon-14 dates and Early Man in the
New World.” In Pleistocene Extinctions: The Search for a
Cause, edited by P. S. Martin and H. E. Wright, 267–286.
New Haven: Yale University Press.
Haynes, C. V., Jr. 1969. “The earliest Americans.” Science 166:
709–715.
Haynes, C. V., Jr. 1970. “Geochronology of man-mammoth sites
and their bearing upon the origin of the Llano complex.” In
Pleistocene and Recent Environments of the Central Plains,
edited by W. E. Dort and E. S. Johnson, 77–92. Lincoln:
University of Nebraska Press.
Haynes, C. V., Jr. 1973. “The Calico site: Artifacts or geofacts.”
Science 181(4097): 305–310.
Haynes, C. V., Jr. 1974. “Archaeological geology of some selected
Paleoindian sites.” In History and Prehistory of the Lubbock
Lake Site, edited by C. C. Black, 133–139. The Museum
Journal XV. Lubbock: West Texas Museum Association.
Haynes, C. V., Jr. 1980. “The Clovis culture.” Canadian Journal of
Anthropology 1(1): 115–121.
Haynes, C. V., Jr. 1982. “Were Clovis progenitors in Beringia?.” In
Paleoecology of Beringia, edited by D. M. Hopkins, J.
V. Matthews, C. E. Schweger, and S. B. Young, 383–398.
New York: Academic Press.
Haynes, C. V., Jr. 1987. “Clovis origins update.” The Kiva 52: 83–
93. Haynes, C. V., Jr. 1991a. “Geoarchaeological and
paleohydrological evidence for a Clovis-age drought in North
America and its
bearing on extinction.” Quaternary Research 35: 435–450.
Haynes, C. V., Jr. 1991b. “More on Meadowcroft radiocarbon
chronology.” The Review of Archaeology 12 (1): 8–14.
Haynes, C. V., Jr. 1993. “Clovis–Folsom geochronology and
climatic change.” In From Kostenki to Clovis: Upper Paleolithic
– Paleo- Indian Adaptations, edited by O. Soffer and N. D.
Praslov,
219–236. New York: Plenum Press.
Haynes, C. V., Jr. 1999. “Monte Verde and the pre-Clovis situation
in America.” Scientific American Discovering Archaeology
Special Report: Monte Verde Revisited, 17–19.
Haynes, C. V., Jr., and G. A. Agogino. 1986. Geochronology
of
Sandia Cave. Contributions to Anthropology 32. Washington
D.C.: Smithsonian Institution Press.
Haynes, G. 1991. Mammoths, Mastodonts, and Elephants:
Biology, Behavior, and the Fossil Record. Cambridge:
Cambridge University Press.
Haynes, G. 2002. The Early Settlement of North America: The Clovis
Era. Cambridge: Cambridge University Press.
Haynes, G. 2009. “Estimates of Clovis-era megafaunal populations
and their extinction risks.” In American Megafaunal
Extinctions at the End of the Pleistocene, edited by
G. Haynes, 39–54. Dordrecht: Springer.
Haynes, G., D. G. Anderson, C. R. Ferring, S. J. Fiedel, D.
J. Grayson, C. V. Haynes, Jr, V. T. Holliday, B. B. Huckell,
M. Kornfeld, D. J. Meltzer, J. Morrow, T. Surovell,
N. M. Waguespack, P. Wigand, and R. M. Yohe. 2007.
“Comment on ‘Redefining the age of Clovis: Implications for
the peopling of the Americas’ by M. Waters & T. Stafford.”
Science 317: 320b.
Haynes, G., and J. M. Hutson 2013. “Clovis-era subsistence:
Regional variability, continental patterning.” In
Paleoamerican Odyssey, edited by K. E. Graf, C. V. Ketron,
and M. R. Waters, 293–309. College Station: Center for the
Study of the First Americans, Texas A&M University.
peopling of North America.” Journal of Taphonomy 8(2–3): 181– Station: Texas A&M University Press.
202. Huckell, B. B. 2014. “But how do we know if it’s Clovis? An exam-
Hedman, W. H. 2010. The Raven Bluff Site: Preliminary Findings ination of Clovis overshot flaking of bifaces and a North
from a Late Pleistocene Site in the Alaskan Arctic. Bureau of Land Dakota cache.” In Clovis Caches: Recent Discoveries and
Management Report, on file at the Alaska State Office of History New Research, edited by B. B. Huckell and J. D. Kilby, 133–
and Archaeology, Anchorage. 152. Albuquerque: University of New Mexico Press.
Hemmings, C. A. 2004. The Organic Clovis: A Single Continent- Wide Ives, J. W., D. Froese, K. Supernant, and G. Yanicki. 2013.
Cultural Adaptation. PhD dissertation, Department of “Vectors, Vestiges, and Valhallas – Rethinking the corridor.”
Anthropology, University of Florida, Gainesville. In Paleoamerican Odyssey, edited by K. E. Graf, C. V.
Hibben, F. C. 1941. Evidence of Early Occupation of Sandia Cave, New Ketron, and M. R. Waters, 149–169. College Station: Center
Mexico, and Other Early Sites in the Sandia-Manzana Region. for the Study of the First Americans, Texas A&M University.
Smithsonian Institution Miscellaneous Collections 99 (23). Jackson, D., C. Méndez, R. Seguel, A. Maldonado, and G. Vargas.
2007. “Initial occupation of the Pacific coast of Chile during
Hill, C. L., and L. B. Davis. 1998. “Stratigraphy, AMS radiocarbon age,
and stable isotope biogeochemistry of the Lindsay Mammoth, late Pleistocene times.” Current Anthropology 48: 725–731.
eastern Montana.” Current Research in the Pleistocene 15: 109– Jackson, L. J. 2006. “Fluted and Fishtail points from southern
112. coastal Chile: New evidence suggesting Clovis- and Folsom-
Hockett, B., and D. L. Jenkins. 2013. “Identifying stone tool cut marks related occupations in southernmost South America.” In
and the pre-Clovis occupation of the Paisley Caves.” American Paleoindian Archaeology: A Hemispheric Perspective, edited
Antiquity 78(4): 762–778. by J. E. Morrow and C. Gnecco, 105–120. Gainesville:
Hoffecker, J. F. 2005. A Prehistory of the North: Human Settlement University Press of Florida.
of the Higher Latitudes. New Brunswick: Rutgers University Press. Jenkins, D. L. 2007. “Distribution and dating of cultural and
Hoffecker, J. F., and S. A. Elias. 2007. Human Ecology of Beringia. paleontological remains at the Paisley 5 Mile Point Caves
New York: Columbia University Press. (35LK3400) in the northern Great Basin: An early
Hoffecker, J. F., W. R. Powers, and T. Goebel. 1993. “The coloniza- tion assessment.” In Paleoindian or Paleoarchaic? Great Basin
of Beringia and the peopling of the New World.” Science 259: 46– Human Ecology at the Pleistocene–Holocene Transition,
53. edited by K. Graf and
Holen, S. R., and K. Holen. 2013. “The Mammoth Steppe hypoth- D. Schmitt, 57–81. Salt Lake City: University of Utah Press.
esis: The Middle Wisconsin (Oxygen Isotope Stage 3) peopling of Jenkins, D. L., L. G. Davis, T. W. Stafford, Jr, P. F. Campos,
North America.” In Paleoamerican Odyssey, edited by T. J. Connolly, L. S. Cummings, M. Hofreiter, B. Hockett,
K. E. Graf, C. V. Ketron, and M. R. Waters, 429–444. K. McDonough, L. Luthe, P. W. O’Grady, K. J. Reinhard,
College Station: Center for the Study of the First Americans, M. E. Swisher, F. White, B. Yates, R. W. Yohe, Jr, C. Yost,
Texas A&M University. and E. Willerslev. 2013. “Geochronology, archaeological
Holmes, C. E. 2001. “Tanana Valley archaeology circa 14,000 to 9000 context, and DNA at the Paisley Caves.” In Paleoamerican
B.P.” Arctic Anthropology 38(2): 154–170. Odyssey, edited by K. E. Graf, C. V. Ketron, and
Holmes, C. E. 2011. “The Beringian and transitional periods in M. R. Waters, 485–510. College Station: Center for the Study
Alaska: Technology of the East Beringian Tradition as viewed from of the First Americans, Texas A&M University.
Swan Point.” In From the Yenisei to the Yukon: Interpreting Lithic Jenkins, D. L., L. G. Davis, T. W. Stafford, Jr, P. F. Campos,
Assemblage Variability in Late Pleistocene/ Early Holocene B. Hockett, G. T. Jones, L. S. Cummings, C. Yost,
Beringia, edited by T. Goebel and I. Buvit, 179–191. College T. J. Connolly, R. M. Yohe, II, S. C. Gibbons, M. Raghavan,

M. Rasmussen, J. L. A. Paijmans, M. Hofreiter, B. M. Kemp, Karr, L. 2015, “Human use and reuse of megafaunal bones in
J. L. Barta, C. Monroe, M. T. P. Gilbert, and E. Willerslev. North America: Bone fracture, taphonomy, and
2012a. “Clovis age Western Stemmed projectile points archaeological interpretation.” Quaternary International
and human coprolites at the Paisley Caves.” Science 337: 301: 332–341.
361 223–228. Kay, M. 2000. “Use-wear analysis.” In The 1999 Excavations at
Jenkins, D. L., L. G. Davis, T. W. Stafford, Jr, P. F. Campos, the
B. Hockett, G. T. Jones, L. S. Cummings, C. Yost, Big Eddy Site (23CE426), edited by N. H. Lopinot, J. H.
T. J. Connolly, R. M. Yohe, II, S. C. Gibbons, M. Ray, and M. D. Conner, 177–220. Southwest Missouri State
Raghavan, University Center for Archaeological Research. Special
M. Rasmussen, J. L. A. Paijmans, M. Hofreiter, B. Publication No. 3, Springfield.
M. Kemp, J. L. Barta, C. Monroe, M. T. P. Gilbert, and Kelly, R. L. 1987. “A comment on the pre-Clovis deposits at
E. Willerslev. 2012b. “Supplementary materials for ‘Clovis Meadowcroft Rockshelter (Letter to the Editor).”
age Western Stemmed projectile points and human coprolites Quaternary Research 27: 332–334.
at the Paisley Caves’.” Science 337 (http://www.sciencemag Kelly, R. L. 2003. “Maybe we do know when people first came
.org/content/337/6091/223/suppl/DC1). to North America; and what does it mean if we do?”
Jennings, T. A., and M. R. Waters. 2014. “Pre-Clovis lithic Quaternary International 109–110: 133–145.
technol- ogy at the Debra L. Friedkin site, Texas: Kelly, R. L., and L. Todd. 1988. “Coming into the country: Early
Comparisons to Clovis Paleoindian hunting and mobility.” American Antiquity
through site-level behavior, technological trait-list, and 53(2): 231–244.
cladistic analyses.” American Antiquity 79(1): 25–44. Kenady, S. M., M. C. Wilson, R. F. Schalk, and R. R.
Johnson, E. 2006. “The taphonomy of mammoth localities in Mierendorf. 2011. “Late Pleistocene butchered Bison
southeastern Wisconsin (USA).” Quaternary International antiquus from Ayer Pond, Orcas Island, Pacific Northwest:
142–143: 58–78. Age confirmation and taphonomy.” Quaternary
Johnson, E. 2007. “Along the ice margin – The cultural taphonomy International 233: 130–141.
of late Pleistocene mammoths in southeastern Wisconsin Kimball, L. R. 1996. “Early Archaic settlement and
(USA).” Quaternary International 169–170: 64–83. technology:
Johnson, M. E. 1997. “Additional research at Cactus Hill: Lessons from Tellico.” In The Paleoindian and Early
Preliminary description of northern Virginia Chapter ASV’s Archaic Southeast, edited by D. G. Anderson and K. F.
1993 and 1995 excavations. Appendix G.” In Sassaman, 149–186. Tuscaloosa: University of Alabama
Archaeological Press.
Investigations of Site 44SX202, Cactus Hill, Sussex County, Krasinski, K. E. 2010. Broken Bones and Cutmarks:
Virginia, edited by J. M. McAvoy and L. D. McAvoy, unnum- Taphonomic
bered pages. Virginia Department of Historic Resources Analyses and Implications for the Peopling of North
Research Report Series No. 8, Richmond. America. PhD dissertation, Department of Anthropology,
Joyce, D. J. 2006. “Chronology and new research on the University of Nevada, Reno.
Schaefer Mammoth (?Mammuthus primigenius), Kenosha Kuehn, S. C., D. G. Froese, P. E. Carrara, F. F. Foit, Jr, N. J.
County, Wisconsin, USA.” Quaternary International 142– Pearce, and P. Rotheisler. 2009. “Major- and trace-element
143: 44–57. characteriz- ation, expanded distribution, and a new
Joyce, D. J. 2013. “Pre-Clovis megafauna butchery sites in the chronology for the latest Pleistocene Glacier Peak tephras in
western Great Lakes region, USA.” In Paleoamerican North America.” Quaternary Research 71: 201–216.
Odyssey, edited by K. Graf, C. V. Ketron, and M. R. Waters, Kunz, M. L. 2010. “Clovis Progenitors: Immigrants or Home
467–483. College Station: Center for the Study of the First Grown?” Abstracts of the SAA 75th Anniversary Meeting,
Americans, Texas A&M University. 15–18 April, St Louis, MO, 144.
Lahaye, C., M. Hernandez, E. Boëda, G. D. Felice, N. Guidon,
S. Hoeltz, A. Lourdeau, M. Pagli, A.-M. Pessis, M. Rasse, by 20,000 BC: The Toca da Peia site, Piauí, Brazil.” Journal
and S. Viana. 2013. “Human occupation in South America of Archaeological Science 40: 2840–2847.
Lavallée, D. 2000. The First South Americans: The Peopling of a
Continent from the Earliest Evidence to High Culture. P. G.
Bahn, trans. (orig. 1993). Salt Lake City: University of Utah
Press.
Lepper, B. T., T. A. Frolking, D. C. Fisher, G. Goldstein, J.
E. Sanger, D. A. Wymer, J. G. Ogden, III, and P. E. Hooge.
1991. “Intestinal contents of a late Pleistocene mastodont
from midcontinental North America.” Quaternary Research
36: 120–125.
Lopinot, N. H., and J. H. Ray. 2000. “Introduction.” In The 1999
Excavations at the Big Eddy Site (23CE426), edited by
N. H. Lopinot, J. H. Ray, and M. D. Conner, 1–16.
Southwest Missouri State University Center for
Archaeological Research. Special Publication No. 3,
Springfield.
Lopinot, N. H., and J. H. Ray. 2007. “Trampling experiments in the
search for the earliest Americans.” American Antiquity 72(4):
771–782.
Lopinot, N. H., J. H. Ray, and M. D. Conner, eds. 1998. The 1997
Excavations at the Big Eddy Site (23CE426) in Southwest
Missouri. Southwest Missouri State University Center for
Archaeological Research. Special Publication No. 2.
Lopinot, N. H., J. H. Ray, and M. D. Conner, eds. 2000. The 1999
Excavations at the Big Eddy Site (23CE426). Southwest
Missouri State University Center for Archaeological
Research. Special Publication No. 3.
Lowery, D. L., M. A. O’Neal, J. S. Wah, D. P. Wagner, and
D. J. Stanford. 2010. “Late Pleistocene upland stratigraphy in
the western Delmarva peninsula, USA.” Quaternary Science
Reviews 29: 1472–1480.
Lynch, T. F. 2001. “On the road again. Reflections on Monte
Verde.” Review of Archaeology 22(1): 39–43.
Mackie, Q., L. Davis, D. Fedje, D. McLaren, and A. Gusick. 2013.
“Locating Pleistocene-age submerged archaeological sites on
the Northwest Coast: Current status of research and future
directions.” In Paleoamerican Odyssey, edited by K. Graf,
C. V. Ketron, and M. R. Waters, 133–147. College Station:
Center for the Study of the First Americans, Texas A&M
University.
MacNeish, R. S. 1971. “Early man in the Andes.” Scientific
American 224: 36–46.
MacNeish, R. S. 2004. “Early inhabitants of the Americas: Pendejo
Cave and beyond.” In Pendejo Cave, edited by R. S. MacNeish
and J. G. Libby, 469–505. Albuquerque: University of New
Mexico Press.
Mashenko, E. V., A. F. Pavlov, V. N. Zenin, S. V. Leshchinskiy, and
L. A. Orlova. 2003. “The Lugovskoe site: Relations between
the mammoth assemblage and Late Palaeolithic man.” In 3rd
International Mammoth Conference, 2003: Program and
Abstracts, edited by J. E. Storer, 77–80. Government of
Yukon Palaeontology Program. Occasional Papers in Earth
Sciences No. 5.
Matisoo-Smith, E., and K. A. Horsburgh. 2012. DNA for
Archaeologists. Walnut Creek: Left Coast Press.
McAvoy, J. M., and L. D. McAvoy. 1997. Archaeological
Investigations of Site 44SX202, Cactus Hill, Sussex County,
Virginia. Virginia Department of Historic Resources Research
Report Series No. 8.
Mead, J. I. 1980. “Is it really that old? A comment about the
Meadowcroft Rockshelter, 1977, an overview.” American
Antiquity 45: 579–582.
Mehl, M. G. 1975. “‘Vertebrate paleomortology at the Cooperton
site.’ In The Cooperton Mammoth: An Early Man Bone
Quarry, edited by A. D. Anderson, 165–168.” Great Plains
Journal 14(2).
Meltzer, D. J. 2009. First Peoples in a New World: Colonizing Ice
Age America. Berkeley: University of California Press.
Meltzer, D. J., J. M. Adovasio, and T. D. Dillehay. 1994. “On a
Pleistocene human occupation at Pedra Furada, Brazil.”
Antiquity 68: 695–714.
Miller, D. S. 2010. Clovis Excavations at Topper 2005–2007:
Examining Site Formation Processes at an Upland Paleoindian
Site along the Middle Savannah River. Occasional Papers 1,
Southeastern Paleoamerican Survey, South Carolina Institute
of Archaeology and Anthropology, University of South
Carolina, Columbia.
Miller, D. S., V. T. Holliday, and J. Bright. 2013. “Clovis across
the continent.” In Paleoamerican Odyssey, edited by K. Graf,
C. V. Ketron, and M. R. Waters, 207–245. College Station: E. Khusnutdinova, S. Litvinov, L. P.
Center for the Study of the First Americans, Texas A&M Osipova, S.
University.
Morrow, J. E., S. J. Fiedel, D. L. Johnson, M. Kornfeld,
M. Rutledge, and W. R. Wood. 2012. “Pre-Clovis in Texas? A
critical assessment of the ‘Buttermilk Creek Complex’.”
Journal of Archaeological Science 39: 3677–3682.
Mulligan, C. J., and A. Kitchen. 2013. “Three-stage colonization
model for the peopling of the Americas.” In Paleoamerican
Odyssey, edited by K. Graf, C. V. Ketron, and M. R. Waters,
171–181. College Station: Center for the Study of the First
Americans, Texas A&M University.
Munyikwa, K., J. K. Feathers, T. M. Rittenour, and
H. K. Shrimpton. 2011. “Constraining the Late Wisconsinan
retreat of the Laurentide Ice Sheet from western Canada
using luminescence ages from postglacial aeolian dunes.”
Quaternary Geochronology 6(4): 407–422.
Nikolskiy, P., and V. Pitulko. 2013. “Evidence from the Yana
Palaeolithic site, arctic Siberia, yields clues to the riddle of
mammoth hunting.” Journal of Archaeological Science 40:
4189–4197.
Overstreet, D. F. 1993. Chesrow, A Paleoindian Complex in
the
Southern Lake Michigan Basin. Case Studies in Great Lakes
Archaeology, Number 2. Milwaukee: Great Lakes
Archaeological Press.
Overstreet, D. F. 1996. “Still more on cultural contexts of
mammoth and mastodon in the southwestern Lake Michigan
basin.” Current Research in the Pleistocene 13: 36–38.
Overstreet, D. F., D. J. Joyce, K. Hallin, and D. Wasion. 1993.
“Cultural contexts of mammoth and mastodon in the
southwest Lake Michigan Basin.” Current Research in the
Pleistocene 10: 75–77.
Overstreet, D. F., D. J. Joyce, K. Hallin, and D. Wasion. 1995.
“More on cultural context of mammoth and mastodon in the
southwestern Lake Michigan basin.” Current Research in the
Pleistocene 12: 40–42.
Overstreet, D. F., and M. F. Kolb. 2003. Guide to Field Trip
to
Southeast Wisconsin Mammoth and Paleoindian Sites. Center
for Archaeological Research, Marquette University Reports
of Investigations No. 03.003, Milwaukee.
Overstreet, D. F., and T. W. Stafford, 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: 70–71.
Perego, U. A., A. Achilli, N. Angerhofer, M. Accetturo, M. Pala,
A. Olivieri, B. Hooshiar Kashani, K. H. Ritchie, R. Scozzari,
Q. P. Kong, N. M. Myres, A. Salas, O. Semino, H.
J. Bandelt, S. R. Woodward, and A. Torroni. 2009.
“Distinctive Paleo-Indian migration routes from Beringia
marked by two rare mtDNA haplogroups.” Current Biology
19(1): 1–8.
Pinson, A. O. 2010. “Paisley Caves, part I: What’s the scoop on the
poop?” Mammoth Trumpet 25(4): 16–20.
Pitulko, V., P. Nikolskiy, A. Basilyan, and E. Pavlova. 2013.
“Human habitation in arctic western Beringia prior to the
LGM.” In Paleoamerican Odyssey, edited by K. Graf,
C. V. Ketron, and M. R. Waters, 13–44. College Station:
Center for the Study of the First Americans, Texas A&M
University Press.
Poinar, H., S. Fiedel, C. E. King, A. M. Devault, K. Bos, M. Kuch,
and R. Debruyne. 2009. “Comment on ‘DNA from pre-Clovis
human coprolites in Oregon, North America’.” Science
325(5937): 148.
Potter, B. A. 2011. “Late Pleistocene and Early Holocene assem-
blage variability in central Alaska.” In From the Yenisei
to the Yukon: Interpreting Lithic Assemblage Variability in
Late Pleistocene/Early Holocene Beringia, edited by
T. Goebel and I. Buvit, 215–233. College Station: Texas
A&M University.
Potter, B. A., C. E. Holmes, and D. R. Yesner. 2013. “Technology
and economy among the earliest prehistoric foragers in
interior eastern Beringia.” In Paleoamerican Odyssey,
edited by
K. Graf, C. V. Ketron, and M. R. Waters, 81–103. College
Station: Center for the Study of the First Americans, Texas
A&M University.
Raghavan, M., P. Skoglund, K. E. Graf, M. Metspalu,
A. Albrechtsen, I. Moltke, S. Rasmussen, T. W. Stafford, Jr.,
L. Orlando, E. Metspalu, M. Karmin, K. Tambets, S.
Rootsi,
R. Mägi, P. Campos, E. Balanovska, O. Balanovsky,
A. Fedorova, M. I. Voevoda, M. DeGiorgio, T. Sicheritz- Ponten, from northern Ohio.” World Archaeology 44(1): 75–101.
S. Brunak, S. Demeshchenko, T. Kivisild, R. Villems, Reimer, P. J., E. Bard, A. Bayliss, J. W. Beck, P. G. Blackwell, C.
R. Nielsen, M. Jakobsson, and E. Willerslev. 2013. “Upper B. Ramsey, C. E. Buck, H. Cheng, R. L. Edwards,
Palaeolithic Siberian genome reveals dual ancestry of Native M. Friedrich, P. M. Grootes, T. P. Guilderson,
Americans.” Nature 505: 87–91. H. Haflidason, I. Hajdas, C. Hatté, T. J. Heaton, D.
Ranere, A., and R. Cooke. 1991. “Paleoindian occupation in Central L. Hoffman, A. G. Hogg, K. A. Hughen, K. F. Kaiser,
American tropics.” In Clovis Origins and Human Adaptations, B. Kromer, S. W. Manning, M. Niu, R. W. Reimer, D.
edited by R. Bonnichsen and K. Fladmark, 237–254. Orono: A. Richards, E. M. Scott, J. R. Southon, R. A. Staff,
Center for the Study of the First Americans, University of C. S. M. Turney, and J. van der Plicht. 2013. “IntCal13 and
Maine. Marine13 radiocarbon age calibration curves, 0–50,000 years
Rasic, J. T. 2011. “Functional variability in the late Pleistocene cal BP.” Radiocarbon 55(4): 1869–1887.
archaeological record of eastern Beringia: A model of late Sanchez, G., V. T. Holliday, J. Arroyo, N. Martínez, and E.
Pleistocene land use and technology from northwest Alaska.” In Gaines. 2013. “El Fin del Mundo, Sonora, Mexico: Where
From the Yenisei to the Yukon: Interpreting Lithic Assemblage Clovis People Hunted Gomphotheres.” Poster presented at
Variability in Late Pleistocene/Early Holocene Beringia, edited by the Paleoamerican Odyssey Conference, Santa Fe, New
T. Goebel and I. Buvit, 128–164. College Station: Texas A&M Mexico, October 17–19.
University Press. Sanchez, G., V. T. Holliday, E. P. Gaines, J. Arroyo-Cabrales,
Rasmussen, M., S. L. Anzick, M. R. Waters, P. Skoglund, N. Martínez-Tagüeña, A. Kowler, T. Lange, G. Hodgins,
M. DiGiorgio, T. W. Stafford, Jr, S. Rasmussen, I. Moltke, S. M. Mentzer, and I. Sanchez-Morales. 2014. “A human
A. Albrechtsen, S. M. Doyle, G. D. Poznik, (Clovis)-gomphothere (Cuvieronius sp.) association ∼13,390
V. Gudmundsdottir, R. Yadav, A.-S. Malaspinas, calibrated yBP in Sonora, Mexico.” Proceedings of the
S. S. V. White, M. E. Allentoft, O. E. Cornejo, K. Tambets, National Academy of Sciences (USA) 111(30): 10972–10977.
A. Eriksson, P. D. Heintzman, M. Karmin, Schaffer, B. S., and B. W. Baker. 1997. “How many epidermal ridges
T. S. Korneliussen, D. J. Meltzer, T. L. Pierre, J. Stenderup, per linear centimeter? Comments on possible pre-Clovis
L. Saag, V. M. Warmuth, M. C. Lopes, R. S. Malhi, human friction skin prints from Pendejo Cave.” American
S. Brunak, T. Sicheritz-Ponten, I. Barnes, M. Collins, Antiquity 62(3): 559–560.
L. Orlando, F. Balloux, A. Manica, R. Gupta, M. Metspalu, Sciulli, P. W. 1982. “Human remains from Meadowcroft
C. D. Bustamante, M. Jakobsson, R. Nielssen, and Rockshelter, Washington County, southwestern
E. Willerslev. 2014. “The genome of a late Pleistocene human from Pennsylvania.” In Meadowcroft: Collected Papers on the
a Clovis burial site in western Montana.” Nature 506: 225–229. Archaeology of Meadowcroft Rockshelter and the Cross
Ray, J. H., N. H. Lopinot, E. R. Hajic, and R. D. Mandel. 2000. Creek Drainage, edited by R. C. Carlisle and J. M.
“Possible pre-Clovis-age artifacts from the Big Eddy Site.” Current Adovasio,
Research in the Pleistocene 17: 68–71. 175–185. Prepared for the symposium “The Meadowcroft
Ray, J. H., N. H. Lopinot, E. R. Hajic, and R. D. Mandel. 1998. “The Rockshelter Rolling Thunder Review: Last Act,” 47th Annual
Big Eddy Site: A multicomponent Paleoindian site on the Ozark Meeting of the Society for American Archaeology,
border, Southwest Missouri.” Plains Anthropologist 43(163): 73– Minneapolis, April 14–17.
81. Sellet, F., J. Donohue, and M. G. Hill. 2009. “The Jim Pitts Site: A
Redmond, B. G., H. G. McDonald, H. J. Greenfield, and M. stratified Paleoindian site in the Black Hills of South Dakota.”
L. Burr. 2012. “New evidence for late Pleistocene human American Antiquity 74(4): 735–758.
exploitation of Jefferson’s Ground Sloth (Megalonyx jeffersonii)

Sherwood, S. C., B. N. Driskell, A. R. Randall, and S. C. Rockshelter Rolling Thunder Review: Last Act,” 47th
Meeks. 2004. “Chronology and stratigraphy at Dust Cave, Annual Meeting of the Society for American Archaeology,
Alabama.” American Antiquity 69(3): 533–554. Minneapolis, April 14–17.
Shott, M. J. 2013. “Human colonization and late Pleistocene lithic Surovell, T. A., and N. M. Waguespack. 2008. “How many
industries of the Americas.” Quaternary International 285: 150–160. elephant kills are 14? Clovis mammoth and mastodon kills
Sistiaga, A., F. Berna, R. Laursen, and P. Goldberg. 2014. “Steroidal in context.” Quaternary International 191: 82–97.
biomarker analysis of a 14,000 years old putative human copro- Surovell, T. A., and N. M. Waguespack. 2009. “Human prey
lite from Paisley Cave, Oregon.” Journal of Archaeological choice
Science 41: 813–817. in the late Pleistocene and its relation to megafaunal
Smallwood, A. M. 2012. “Clovis technology and settlement in the extinc- tions.” In American Megafaunal Extinctions at the End of
American southeast: Using biface analysis to evaluate the Pleistocene, edited by G. Haynes, 77–105. Dordrecht:
dispersal models.” American Antiquity 77(4): 689–713. Springer. ThisIsBozeman.com [website] [1967]. “Digging the
Smallwood, A. M., D. S. Miller, and D. Sain. 2013. “Topper site, Lindsay Mammoth.”
South Carolina: An overview of the Clovis lithic assemblage http://www.thisisbozeman.com/detail/digging-
from the Topper Hillside.” In In the Eastern Fluted Point the-lindsey-mammoth-1967/, accessed 23 June
Tradition, edited by J. A. M. Gingerich, 280–298. Salt Lake 2014.
City: University of Utah Press. Vasil’ev, S. A. 2011. “The earliest Alaskan archaeological
Smith, H. L., J. T. Rasic, and T. Goebel. 2013. “Biface traditions of record: A view from Siberia.” In From the Yenisei to the
northern Alaska and their role in the peopling of the Yukon: Interpreting Lithic Assemblage Variability in Late
Americas.” In Paleoamerican Odyssey, edited by K. Graf, C. Pleistocene/ Early Holocene Beringia, edited by T. Goebel
V. Ketron, and and I. Buvit, 119–127. College Station: Texas A&M
M. R. Waters, 105–123. College Station: Center for the Study University Press.
of the First Americans, Texas A&M University. Waguespack, N. M. 2007. “Why we’re still arguing about the
Stanford, D. 1982. “A critical review of archaeological evidence Pleistocene occupation of the Americas.” Evolutionary
relating to the antiquity of human occupation of the New Anthropology 16: 63–74.
World.” In Plains Indian Studies, edited by D. H. Ubelaker Waguespack, N. M., and T. A. Surovell. 2003. “Clovis hunting
and H. J. Viola, 202–218. Smithsonian Contributions to strat- egies, or how to make out on plentiful resources.”
Anthropology No. 30, Washington D.C. American Antiquity 68(2): 333–352.
Stanford, D. J., and B. A. Bradley. 2012. Across Atlantic Ice: The Waters, M. R., S. L. Forman, T. A. Jennings, L. C.
Origin of America‘s Clovis Culture. Berkeley: University of Nordt,
California Press. S. G. Driese, J. M. Feinberg, J. L. Keene, J.
Stevens, D. E., and D. A. Agogino. 1975. Sandia Cave: A Study in Halligan,
Controversy. Eastern New Mexico University. Contribution to
Anthropology 7 (1), Portales, New Mexico.
Stuckenrath, R., J. M. Adovasio, J. Dohanue, and R. C. Carlisle.
1982. “The stratigraphy, cultural features and chronology at
Meadowcroft Rockshelter, Washington County, southwestern
Pennsylvania.” In Meadowcroft: Collected Papers on the
Archaeology of Meadowcroft Rockshelter and the Cross
Creek
Drainage, edited by R. C. Carlisle and J. M. Adovasio, 69–
90. Prepared for the symposium “The Meadowcroft
A. Lindquist, J. Pierson, C. T. Hallmark, M. B. Collins, and of Chicago Press.
J. E. Wiederhold. 2011a. “The Buttermilk Creek Complex and West, F. H. 1996. American Beginnings: The Prehistory and
the origins of Clovis at the Debra L. Friedkin Site, Texas.” Palaeoecology of Beringia. Chicago: University of Chicago
Science 331: 1599–1603, and Supporting Online Material DOI: Press.
10.1126/science.1201855. West, F. H. 1999. “The inscrutable Monte Verde.” Scientific
Waters, M. R., S. L. Forman, T. W. Stafford, Jr, and J. Foss. 2009. American Discovering Archaeology Special Report: Monte
“Geoarchaeological investigations at the Topper and Big Pine Verde Revisited 15–16.
Tree sites, Allendale County, South Carolina.” Journal of Wheat, A. D. 2012. “Survey of professional opinions regarding the
Archaeological Science 36: 1300–1311. peopling of the Americas.” The SAA Archaeological Record
M. R. Waters, C. D. Pevny, and D. Carlson, eds. 2011b. Clovis Lithic 12 (2): 10–14.
Technology: Investigations at a Stratified Workshop at the Gault Willig, J. A., and C. M. Aikens. 1988. “The Clovis-Archaic
Site. College Station: Texas A&M University Press. interface in Far Western North America.” In Early Human
Waters, M. R., and T. W. Stafford, Jr. 2007. “Redefining the age of Occupation in Far Western North America: The Clovis-
Clovis: Implications for the peopling of the Americas.” Science Archaic
315: 1122–1126. Interface, edited by J. A. Willig, C. M. Aikens, and
J. L. Fagan, 1–40. Nevada State Museum. Anthropological
Waters, M. R., and T. W. Stafford, Jr. 2013. “The first Americans: A Papers No. 21, Carson City.
review of the evidence for the late-Pleistocene peopling of the
Yesner, D. R., C. M. Barton, G. A. Clark, and G. A. Pearson. 2004.
Americas.” In Paleoamerican Odyssey, edited by K. Graf, “Peopling of the Americas and continental colonization: A
C. V. Ketron, and M. R. Waters, 541–560. College Station: mil- lennial perspective.” In The Settlement of the American
Center for the Study of the First Americans, Texas A&M
Continents: A Multidisciplinary Approach to Human
University.
Biogeography, edited by C. M. Barton, G. A. Clark,
Waters, M. R., and T. W. Stafford, Jr. 2014. “Redating the Mill Iron
site, Montana: A reexamination of Goshen Complex chronol- D. R. Yesner, and G. A. Pearson, 196–213. Tucson:
University of Arizona Press.
ogy.” American Antiquity 79(3): 541–548.
Waters, M. R., T. W. Stafford, Jr, H. G. McDonald, C. Gustafson, Young, C., and S. Gilbert-Young. 2007. “A fluted projectile point
base from Bering Land Bridge National Preserve.” Current
M. Rasmussen, E. Cappellini, J. V. Olsen, D. Szklarczyk, Research in the Pleistocene 24: 154–156.
L. Juhl Jensen, M. T. Gilbert, and E. Willerslev. 2011c. “Pre- Zenin, V. N., E. N. Maschenko, S. V. Leshchinskiy, A. F. Pavlov,
Clovis mastodon hunting 13,800 years ago at the Manis Site,
Washington.” Science 334: 351–353. P. M. Grootes, and M.-J. Nadeau. 2003. “The first direct evi-
dence of mammoth hunting in Asia (Lugovskoye site, western
Webb, S. D. 2006. “Mastodon tusk recovery.” In First Floridians and Siberia).” In 3rd International Mammoth Conference, 2003:
Last Mastodons: The Page-Ladsen Site in the Aucilla River,
Program and Abstracts, edited by J. E. Storer, 152–155.
edited by S. D. Webb, 333–342. Dordrecht: Springer.
Government of Yukon Palaeontology Program. Occasional
West, F. H. 1981. The Archaeology of Beringia. Chicago: University
Papers in Earth Sciences No. 5, Whitehorse.

Author's Biography

Gary Haynes earned his PhD from Catholic University of America in 1981. He is Foundation Professor of
Anthropology at the University of Nevada, Reno, where he has worked for 30 years. His research interests are
North America’s first people, taphonomic and actualistic studies of large mammals, and southern African
prehistory.

You might also like