Next Article in Journal
Last Interglacial Climate in Northern Sweden—Insights from a Speleothem Record
Next Article in Special Issue
Sparse Radiocarbon Data Confound Culture-Climate Links in Late Pre-Columbian Amazonia
Previous Article in Journal
Approach for Analysis of Land-Cover Changes and Their Impact on Flooding Regime
Previous Article in Special Issue
Setting the Stage: The Late Pleistocene Colonization of North America
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Paleoamerican Occupation, Stone Tools from the Cueva del Medio, and Considerations for the Late Pleistocene Archaeology in Southern South America

CONICET-IGEBA, Laboratory of Geophysics “Daniel A. Valencio”, Department of Geological Sciences, FCEN, UBA, Ciudad Autónoma de Buenos Aires C1428EGA, Argentina
Submission received: 13 June 2019 / Revised: 16 July 2019 / Accepted: 5 August 2019 / Published: 12 August 2019

Abstract

:
Archaeological excavations at the Cueva del Medio performed during the 1980s and 1990s yielded an important record of both faunal and stone tool remains, as well as data, to discuss issues that occurred during the Terminal Pleistocene. Due to that, the shaped Paleoamerican artifacts collected in the author’s excavations were partially informed. The present article provides unpublished data on the field-work, the results of a techno-morphological analysis of the stone tools, and considerations about early hunter-gatherer societies along with their regional paleo-environmental interactions, as well other topics regarding the regional archaeological process during the last millennium of the Pleistocene. Findings from there have been extremely useful for discussing diverse paleo-ecological and archaeological topics and have extended the knowledge and discussions about different Pleistocene scientific issues, mainly related with flora, fauna, and the colonization of southern Patagonia.

1. Introduction

During the last millennium of the Pleistocene at ~11–10 kya (11,000–10,000 uncalibrated radiocarbon years before present (BP)) or ~13–12 cal. kya (13,000–12,000 calibrated radiocarbon years BP), an unquestionable fact is that the New World was populated from northern North America to the far south of South America [1,2]. At that time, the “fishtail”, “Fell’s cave”, or simply “Fell” projectile point (FP) had an extraordinary dispersion throughout Central and South America and was becoming a widespread Paleoamerican (PA) marker [2].
Despite isolated finds performed during the 19th century [2], the first FPs in stratigraphic contexts were performed in the southern tip of Patagonia, which is a large geographic region of ~1,000,000 m2 shared by the Republics of Argentina and Chile in southern South America. In the history of the PA studies, this region is well known from the pioneering work in the 1930s of Junius Bird, whose discoveries linked the archaeology of Patagonia with the New World colonization process. Just a few years after the revolutionary discoveries at the Blackwater Draw, Folsom, and Lindenmeier sites demonstrating associated finds of Clovis and Folsom points with remains of Ice-Age fauna in the North American Great Plains, nearby the Magellan straight in southern Chile, Junius Bird excavated the Pali Aike and Fell’s caves. There, in the lower stratigraphic layers, he found the distinctive FP in context with the remains of extinct fauna [3,4,5]. However, since those pioneering discoveries showing that hunter-gatherers using FPs were contemporaneous with the Pleistocene fauna, except for a doubtful claimed PA site excavated during the 1970s [6], no reliable evidence of PA occupations had been found until the 1980s. In fact, after half a century without any clear findings corroborating his discoveries, the southern tip of South America fortunately saw the discovery and excavation of several sites that confirmed the exploitation and consumption of extinct fauna by hunter-gatherer societies. Thus, the Magellan Strait and its surroundings provided a series of previously unthinkable findings related to the oldest human occupations in the Americas [7,8,9]. At that time, because of a lack of reliable evidence recovered through careful excavations, the technology and subsistence of the region’s oldest human populations were subjected to diverse conjectures and interpretations (e.g., [10,11,12]).
The archaeological excavations at the Cueva del Medio (CM) that were performed during the 1980s and 1990s yielded an important record of both faunal and stone tool remains, as well as data, to discuss the anthropological, paleontological, and paleo-environmental issues that occurred during that time-span. Accordingly, along with the findings performed at the Tres Arroyos rock-shelter and Lago Sofía Cave [7,9], the CM became one of the first links in a chain of findings that allowed an approach to be established with new perspectives and reliable information on the initial regional settlement. As a result of the aforementioned excavations and recovered vestiges, an important number of investigations were carried out that shed light on very diverse regional chronological, environmental, and cultural topics (e.g., [8,11,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31]), among others.
Because the shaped PA artifacts collected in the author’s excavations were partially informed [13,14,17], the present article provides first-hand unpublished data on the field-work, a techno-morphological analysis of the stone tools, as well as considerations about the early hunter-gatherer societies in southern Patagonia.

2. Research Activities at the Cueva Del Medio

2.1. Location and Historical Brief

Shared by Chile and Argentina in the southern Andean Cordillera (Figure 1), the Southern Patagonian Ice Field is one of two remnants of the Patagonian Ice Sheet, which covered the whole of southern Chile during the last glacial period [32]. On its south-western slope, the orography is characterized by mountain ranges (e.g., Monumento Moore, Señoret, Chacabuco, and Prat) with an average height of ~1500 m asl (Figure 2). Belonging to the Cerro Toro geological complex [33], there is situated the Cerro Benítez (CB), a small hill in Ultima Esperanza Province, Magallanes Region, southern Chile. Situated on its southern flank, the Cueva del Mylodon Natural Monument protects the world-famous Mylodon cave [34,35,36] and other rock shelters of significant size that yielded an important paleontological and archaeological record (Figure 3).
According to regional historian Mateo Martinic [35], at the beginning of 1895, this locale aroused worldwide scientific interest because in the Mylodon cave, a thick piece of hide with embedded dermal ossicles and long reddish hairs was found in an excellent state of conservation, but not associated with any known living animal. Further visits to the area were made in the next two years, and more bones, skin, and dung were found that were assumed to be the same unknown species. It was identified as a ground sloth, named the mylodon, an animal related to other extinct Pleistocene herbivorous mammals. However, there was some speculation that it was still living in some places in Andean Patagonia [37]. Among the scientists who visited the Benitez hill was Rodolfo Hauthal, a German researcher, who at that time was working at the Museo de La Plata, Argentina. Among other activities performed on the CB, with scientific aims, he made the first excavations at the CM [38]. At the same time, the Daily Express in London organized an expedition with the aim of finding some living exemplars of this mega-mammal. Then, during the period spanning 1899–1902, like the other caves in the hill and also looking for the paleontological remains, the CM was visited by relic hunters who destroyed a significant portion of the cave deposits. In this way, a large number of paleontological specimens were recovered and mostly sold to European museums.
The CM (51°35′ S, 71°38′ W) is situated ~1 km from the Mylodon cave in the Mylodon Cave National Park (Figure 3). It is a large cave ~90 m long by ~50 m wide at its entrance (Figure 4 and Figure 5), which got its name (Cave of the middle) because it is located between the Mylodon cave and the Chica cave, which is the smallest significant shelter on the CB, also, it is medium in size between both of these caves. For most of the 20th century, the CM went into oblivion until the beginning of the 1980s, when Prieto (at that time a graduate student of philosophy and a team member of the IP archaeologist Massone), together with his sister Mónica and friend Romero, re-discovered the cave, whose location, visibility, and accessibility were unknown. Hence, on his behalf, I was invited to perform excavations in Ultima Esperanza by the Instituto de la Patagonia (IP) founder and director, M. Martinic. The field-works were performed between 1986 and 1993 under the auspices and institutional support of the IP, in conjunction with the Universidad de Magallanes (UM). As seen in the respective section below, funding for the field-work and laboratory research were granted to the IP and the author of this paper by different international institutions.
It is worth remembering that at that time, due to a lack of a staff archaeologist and with few authors contributing to the IP publication, the writer of this article collaborated extensively in good faith for its continuity as an institution. According to what can be observed in numerous newspaper articles, at that time the CM finds had a significant impact on the local community. Then, fortunately, after their excavations and discoveries, the IP and some members of the excavation team were incorporated into the UM staff.
Mainly with paleontological and stratigraphic aims, new excavations were recently performed in 2011 and 2012 by deepening the previously excavated area, and an additional small sector close to the cave’s western wall [29].

2.2. Excavations and Findings

Due to its size and for purely descriptive and analytical purposes, the cave was divided into three portions as follows: the entrance, the interior and the back [15]. They are respectively located in its mouth sector and drip line, and around and behind the large and important conglomerate blocks that had fallen from the cave walls and ceiling in the center (Figure 4c,d and Figure 5). As mentioned previously, the site was widely disturbed by looters and the initial excavations performed at the turn of the 20th century, which left a lot of holes (e.g., Figure 5), hence, in diverse sections of the cave’s sedimentary infilling, important accumulations of sediments had been removed and formed. To investigate the archaeological and paleontological remains that were still potentially present, they were carefully screened and a number of remains were recovered. After this removal, it was observed that covered by them, in the front, there were several intact sections of deposits. Also, surrounding the central large blocks, a layer of pebbles that had detached from them had sealed the sediments, making them impossible to have been excavated by the looters. Further, a small section at the end of the cave was also undamaged. Then, the excavations were performed in the deposits in these sectors, mainly in the center (Figure 6). The grids were planned, bearing in mind a north-south axis, and numbered from the cave’s mouth toward the back and from east to west. All of the findings were mapped and recorded.
With regard to its stratigraphy, the published reports on the CM have only made observations based on the first three expeditions. Furthermore, the excavations carried out in the cave’s different sectors revealed a more complex stratigraphy that was formed by sediments of exogene and endogene origin [39,40]. At the entrance and in the central section, the sediments are composed mainly of sand and cobbles originating from the conglomerates that formed the cave. In that area, the stratigraphic sequence containing the archaeological record showed four strata, as follows: the first is a fine a gray powdery surface level. The second is below the former in the eastern part, it can be observed from the surface towards the west and is composed of clasts and pebbles of various sizes mixed with fragments of conglomerates that originated from the wall and ceiling, in some places this level is highly compacted and acted as a seal for the PA archaeological level (Figure 7 and Figure 8). Underlying this stratum and still having some pebbles, there is a sandy reddish/yellowish level that is not present in the totally excavated surface (Figure 7), hence, in a large part of the excavation the stratum of pebbles and clasts of different sizes overlies layer four. Under a gray sandy layer and below a humid clayed deposit with rounded clasts and pebbles, layer five was identified. Below the archaeological excavations, a redeposited volcanic ash layer from the Reclús volcano that originated in the R1 eruption, which occurred at ~12.6 kya was recently identified [29]. Different from the pre-Holocene deposit, the sedimentary section containing the archaeological remains witnessed a very low deposition rate and non-conformities due to episodic sdimentation [41]. Then, the ~50 cm deposit registered a discontinuous sedimentary and archaeological record with gaps since the last millennium of the Pleistocene and Holocene. Originating in this way, there is a mix of diverse kinds of archaeological and paleontological materials in some sectors e.g. [14,42].
The excavations allowed the discovery of a remarkable archaeological level restricted to the excavated area and performed at the front near the entrance, but mainly in the interior in the central section. Because of that, the highly disturbed surface of the cave did not allow us to test whether there was any continuity between the occupation of the frontal and central sectors. It was not possible to know whether there was a specific continuous spatial occupation of the cave. Man-made remains were also found in the grids located behind the blocks underlying a surficial calcareous crust and dated at ~4.2 kya [16], which agrees with similar dates obtained later on [39]. Following the natural slope of the cave’s infilling, the PA record overlies layer four, thus, it was only excavated up to its upper part. Due to the dip in the deposit, the measurements from the datum varied in depth from around −20/25 to −125/130 cm according to the excavated sector (Figure 7), as expected.
In a broad sector in the interior, the PA remains were covered by a thick layer of pebbles and cobbles, made up of a disaggregation of blocks which had fallen from the cave walls and ceiling (Figure 7, Figure 8 and Figure 9). In general, the PA is a widespread archaeological level, suggesting that early hunter-gatherers used a significant portion of the cave, mostly at its center, but also close to the entrance. This level provided a lot of information on the paleoenvironment and the PA way of life. Due to the existence of several hearths, broken bones, red ochre, and various kinds of stone and bone implements (e.g., flaker, tube, and engraved artifact), the CM could be considered a multiple activity area site (e.g., [14,19]) formed by short-term occupations in different places in the cave.
Not comparable with bone accumulations that exist in several North American PA sites (e.g., [43,44]), among others, but like other “bone beds” from the southern cone (e.g., [45]), showing diverse clusters of remains, the PA record in the CM is a relatively thin level that follows the natural sediment dip. It shows different degrees of reliability according to its location in the excavation. In several sectors (e.g., at the front), but mostly in those covered by rocks, it was undisturbed (Figure 9 and Figure 10) suggesting a good integrity of the record [31]. Other places showed a palimpsest, or a mix of archaeological materials derived from a variety of animal or human events or actions [46] (p. 9). In fact, in some grids there is a mix of natural (e.g., unmodified Mylodon vertebrae, see Figure 1 of [16] and cultural (burn and calcined specimens, or with cut marks) bones (sensu [47]). As observed during the field-work and demonstrated by several taphonomic observations on the bones recovered from the excavations, as well from the surface and sediment accumulations caused by the looters’ activities, many of them were affected by diverse natural and cultural modifications [11,14,22,28,31,48], among others.
Among the most significant finds, a lithic assemblage consisting of diverse unifacial and bifacial artifacts associated with extinct and extant fauna was recovered (e.g., [14,16,22,23]). Faunal remains exhumed at the CM include Felis onça, Mylodon (?) listai., Dusicyum avus, Smilodon, Lycalopex culpaeus, horse (Hippidion saldiasi), Lama gracilis, Lama guanicoe, and a bigger camelid, initially classified as Lama morphotype cf. L. owenii [20,21,22,23,29,48]. Currently, the latter is considered to be a Camelidae [31] and is present in several archaeological and paleontological sites along the Chilean Cordillera [49]. The initial investigations at the CM showed that during the late Pleistocene there was faunal diversity in southern Patagonia [17,21,22,23], and that the earliest hunter-gatherers exploited some of the species of this fauna (Figure 8, Figure 9 and Figure 10). As observed during the archaeological excavations and by bone analysis, its largest accumulation comes from layer four, mostly Camelids, but also Hippidion saldiasi, were consumed by the earliest hunter-gatherers [14,17,21,22,23,31] (Figure 10), Numerous dermal ossicles, several vertebrae and other elements from the Mylodon were exhumed at the top of layer four, mainly in the NW sector of the main excavation (grids 27/11, 28/10 to 28/13, 29/12, and 29/13). Nevertheless, there is no evidence of its consumption [31,50].
Different opinions have been expressed about the status of extinct fauna in the consumption behavior of the earliest regional inhabitants (e.g., [6,11,12,51]), and in particular about the exploitation of Hippidion saldiasi [52]. However, it is currently accepted that hunter-gatherers living during the Pleistocene/Holocene transition had a generalized diet with regional variations, using extant and extinct fauna as a resource [2,7,8,9,17,18,19,20,21,22,31,50,53]. In the southern cone in particular, current research on early sites reveals that a wide range of animals were consumed by humans living at ~11–10 kya [53,54,55,56,57,58,59], among others. The above-mentioned fauna were not only exploited for food consumption, their bones were also used as raw materials for making a variety of goods [60].

2.3. Chronological Data

At the CM, more than 50 radiocarbon dates on diverse paleontological and archaeological materials were performed [8,14,16,24,35,39,61,62]. The samples for radiocarbon analysis came from the archaeological excavations as well bone specimens from the looter’s pits. Except the three samples coming from the apparently mixed deposit containing stemless triangular points and extinct fauna, the remaining ones all belongs from the PA level. A detailed list of radiocarbon assays spanning the oldest human occupations is shown in Table 1. It depicts the calibrated ages calculated using the IntCal13 calibration program [63]. In addition, Figure 11 shows the calibrated range of those corresponding to the Late Pleistocene ages. As can be seen in both depictions, except for the two younger dates of ~9.5 and 9.7 kya, the assays related to the PA occupations belong to the last millennium of the Pleistocene range between ~10.3 and 11.0 kya, and the highest number of occurrences varies between ~10.3 and 10.8 kya. The ~9.5 and 9.7 kya dates from the FP level may be considered outsiders from the time span when the FP makers lived [42,53,63,64,65,66,67].
As mentioned previously, despite the archaeological remains with triangular projectile points and their corresponding lithic assemblage being mostly present in layer three, there is a sector in the transition between layers three and four with a mix of triangular points and Pleistocene fauna (see Figure 14 in [8]). Even a few accelerator mass spectometry (AMS) dates from the bone samples from that level yielded similar ages to those obtained from the FP occupation [24].

3. Paleo American Stone Tools at the CM

The stone tools found at the CM came from the surface, old diggings, and the excavations described above. In the central sector, the archaeological finds were as follows: layer two yielded scattered artifacts, a probable early/middle Holocene lithic assemblage with triangular stemless points in layer three, finally, in the upper part of level four were the PA remains at varying depths as they followed the natural dip in the sedimentary infill (Table 2).
Despite some having been previously analyzed, documented, and reported [8,13,18,26] in Figure 3, most of the PA pieces recovered following the third expedition were not reported in detail. Including both bifacial and unifacial flaked implements, the specimens (n = 31) described here came from the grids mainly situated in the central part (Figure 12), principally from the non-mixed sectors covered by the pebbles and blocks that had fallen from the walls of the cave. Pieces that were only considered to be of doubtful PA origin were not accounted for in this article. The collection was analyzed according to the guidelines and record sheet proposed by Aschero [68]. The salient information regarding origin, dimensions, raw materials, tool class, and other salient information is depicted in Table 2. The most notable lithic artifacts are illustrated in Figure 13, Figure 14, Figure 15 and Figure 16.
Most of the rocks used for making the PA tools were acquired in the secondary sources that were locally available in the glacial deposits surrounding the CB [8,69]. Despite a lack of scientific provenance analysis studies, it has been suggested that the chalcedony came from the Sierra de Los Baguales, a place located ~108 km north of the site [50]. With regard to the first step in the reduction sequences, throughout the excavated sectors, no evidence of blank obtaining activity was detected. As previously reported up to the second expedition Figures 19 and 20 of [8] and as observed in the following field-work, the lithic debitage suggests that shaping and re-sharpening activities were performed. Then, bearing in mind the evidence from the excavation, like their other survival exploitation activities (e.g., wood for fuel, hunting), the PA groups might have obtained the flakes off site.
During the former field-work, two FPs were exhumed (Figure 13a,b and Figure 14a,b), furthermore, an additional broken specimen was recovered [29]. The initial finds were similar in shape with the expected FP “typical” variations previously found in the Fell and Pali Aike caves [3,4,5]. However, a more recent find has a slight difference in its stem and shoulder (Figure 13c), mainly because in its base it lacks the ears that define “typical” FPs. In this way, it constitutes one of the variations, namely, the contracting stem observed in other South American localities (see Figure 21 of [2]), some at the Jaywamachay rock shelter (see Figure 2c of [66]), El Inga, Ecuador (see Figure 22c of [70]), and Paso Calderón, Uruguay (see Figure 2 of [71]). Interestingly, at that time it was not recognized as an FP, and almost at the same time as the initial systematic work at CM, a similar specimen was found on the surface at one of the surveyed sites in the Epullán locality, NW Patagonia [72]. Coincidently, in that area, the Epullán cave yielded an archaeological record that spanned the FP makers [19,73,74].
The FPs exhumed in the previous diggings seem to be re-sharpened artifacts, especially one with a clear flute-like impact fracture originating from its tip (Figure 13b and Figure 14b). In this regard, some doubts remain about the smaller one (Figure 13a and Figure 14a), which was probably rejuvenated close to its tip. With a narrow blade, this piece might have been one the lanceolate variants made by hunter-gatherers using this sort of point (see Figure 22 of [2] and Figure 19b of [70]), [21,75] Figure 5d,o. The fracture in the stem as illustrated in Figure 13c was probably produced by an impact [8,76]. As can be seen in Figure 13b,c, from a manufacturing viewpoint the FPs were finished with short irregular retouches achieved by pressure flaking not very carefully applied to a flake blank, a kind of pattern regularly used by FP makers [69,75,77,78,79]. Like other FPs, the specimens illustrated in Figure 13a,b show abrasions on the edges of the stems (e.g., [53,63,70,77]).
With the exception of the projectile points and a few fractured bifacial specimens, one possibly broken during the early manufacturing stage (Figure 16b) made by bifacial reduction, the remaining units (n = 27) were unifacial, mainly side scrapers and knives of various dimensions. However, the remarkable ones are those that are larger in size (Table 2, Figure 14c,d and Figure 16c,d). In Argentinean lithic studies, “knives” and “side scrapers” are arbitrarily differentiated by the value of the edge angle cf. [80], however, this sort of implement might have had diverse functions, which is why some archaeologists call them “multipurpose tools” [81]. For example, actualistic experimental studies demonstrated that “side scrapers” were highly useful for sawing whale bones [82]. Apparently, the largest unifacials might have not been subjected to repetitive re-sharpening, as they were habitually subjected [83,84]. They were probably shaped and employed at, or near the site, and then left there after use. Bearing in mind the initial distinction between “curated” and “expedient technologies” [85] and other discussions (e.g., [86]), it could be suggested that some of those from the CM might have been curated tools. In fact, as “techno-units” of more complex utensils [87] (p. 8), FPs were part of “personal gear, while large-sized unifacials seem mostly to have been related to a curated behavior for furnishing the site [85].
The unifacial implements were made using tertiary flake blanks obtained by hard hammer, direct percussion flaking, mostly exhibiting flat and diffuse force application bulbs, as experimentally observed in the tuff from the CB [88], the retouches vary from short parallel to scaliform. When a flake’s dorsal and ventral faces are visible, most retouches are direct (Figure 15c,d and Figure 16), and inverse in a few cases (Figure 15a,b). It is highly probable that final shaping was performed by simply retouching a blank’s edges using the same technique, but with another variant and/or holding position, and using soft or semi-soft percussors (e.g., [89]). However, as can be seen in Figure 16c, there is a specimen that shows a more complex manufacture that consisted of flaking the blank, with the covering unifacial flaking likely employing soft percussion, after which it was finished by retouching the edges. A comparable manufacture was observed in coeval assemblages found at the El Ceibo and Los Toldos caves in the Deseado massif in Santa Cruz Province, Argentina (see Figure 7b of [2] and see Figure 91 of [90,91]). Interestingly, a comparable production was recently identified as being early Patagonian Holocene lithic assemblages [89,92,93]. It is worth mentioning that in locations where local raw materials were available, large unifacial tools are present at sites where there is evidence of hunter-gatherers living during the last millennium of the Pleistocene (e.g., see Figure 6 of [63], [94,95]), but also during the middle and late Holocene occupations in southern Patagonia [3,4,5,89,92,93].
Morphologically, the largest unifacial instruments are akin to the so-called “level 11 industry,” which, without FPs, were found in the lower levels of the El Ceibo and Los Toldos caves, but with a similar chronology [90,91]. There were also some with FPs at the Pali Aike and Fell caves [5] (Figure 16f–g) as well as at the Piedra Museo rock shelter, see (Figure 8.24–8.29 of [95]). Further, the lithic finds at the CM and other sites allowed us to understand locales where there were no FPs, but with other comparable artifacts, earlier attributed to “unifacial flake industries” (e.g., [12,18,91]), see pp. 170–171 of [63]. Advances in research show that several sites without FPs but with unifacial tools (e.g., Cueva del Lago Sofía, e.g., [9,96,97]) may have been part of these sites’ variability, where there were hunter-gatherers who used these kinds of points [19,63,98]. A similar fact might be expected for “unifacial traditions,” with large utensils proposed in other South American countries such as Brazil [99,100], where an important FP surficial record was recently reported [101,102]. Interestingly, a specimen from Bahia State (see Figure 22e of [2], Figure Ik of [78], and Figure 1c of [103]) was found along with large unifacial side scrapers made from the same material [104,105].

4. The Early Inhabitants of the Cueva del Medio and Considerations on Late Pleistocene Archaeology in Southern Patagonia

As a part of the significant changes that were occurring on Earth during the terminal Pleistocene, South America showed a different vegetation and environmental picture than today (Figure 17). Several studies have already been undertaken to provide insights into the paleoenvironment prevailing at the time of the PA occupation. In fact, the studied area was characterized by a periglacial environment with a tundra biome [53,106,107,108]. The palynological sampling performed during the CM excavations yielded a pollen and spore diagram that included the late-glacial-Holocene transition to the Late Holocene (Figure S1 in Supplementary Materials). The samples taken from the top of layer four at the PA level show a predominance of Gramineae (grass), following an episode of treeless grass steppe, as observed in levels two and three, while Nothofagus (southern beech) in the sample from layer one implies the late arrival of trees at the CM [26]. The archaeological evidence demonstrates that species of this flora were employed by early hunter-gatherers; in fact, andragogical studies on burned wood from PA hearths show that they used Nothophagus pumilio and, probably, Empetrum rubrum [8]. Bearing in mind the local paleontological record, this sort of vegetation was consumed by different species of herbivorous animals such as ground sloths (Mylodon sp.), horses (Hippidion saldiasi), camelids (Lama cf. owennii, Lama gracilis, Lama guanicoe), and cervids (Hippocamelus bisulcus), some of which provided different kinds of resources, mainly food and raw materials for making bone implements [60] and, probably, other goods.
A significant climatic feature of southern Patagonia is the westerlies, which are the prevailing winds in the middle latitudes of the Earth’s atmosphere, blowing from west to east between the high-pressure areas of the subtropics and the low-pressure zones over the poles [110,111], a phenomenon that might have had a certain influence on southern Patagonian prehistory [112,113]. During the PA period, regional glaciers were retreating and the area under study was still a harsh environment [1]. Despite this, the archaeological and paleo-ecological evidence shows that this landscape provided a favorable habitat when the earliest foragers arrived at the CB. During the earliest periods, they were possibly not entirely familiar with the surroundings, however, as time progressed, they undoubtedly became more knowledgeable. It has been suggested that it takes considerable time, energy, and experimentation to learn about and know a landscape in-depth [114], so mobility patterns would have needed to be adjusted in order to exploit the available survival supplies more effectively. However, for basic survival resources such as rocks, wood fuel, and some animals, their familiarization with their habitat may have been rapid [63,114]. In this regard, the ecosystem offered significant resources for existence, as well as rock shelters for refuge, mainly when adverse weather made outdoor life difficult.
Despite the harsh environment in the CB at the time of the PA occupation at the CM, this cave offered an excellent shelter for humans, especially during the winter. In order to control the CM’s advantage as a refuge, during the fifth and sixth expeditions the temperature of the CM was systematically taken three times a day in the portions mentioned above, as well as outside the cave. As can be observed in Figure S2 of Supplementary Materials, both in the central part and at the rear, temperatures ranging from 8.5–9.5 °C were relatively constant, while those at the front were ~1–2 °C higher, but more variable and related to the outside values, which were generally ~5 to 10 °C higher during the summer. Despite some adverse atmospheric phenomena occurring outside-of course, during the time of the PA occupation these values may have been lower-the CM offers excellent protection, mainly in the central portion, where the influence of what is happening outside is less, especially wind and rain. Like other caves that this author has excavated and lived in southern Patagonia e.g., [115,116,117], among others, many offer excellent protection for humans and animals. In addition, they also protect from the large and unpredictable daily local climatic variations, because on the same day it can rain, there can be strong winds, and the sun can shine.
It is worth mentioning that the majority of the taxa found at the CB represent species that were affected by the large worldwide Late Pleistocene mass extinction, a highly complex topic in the Earth’s history (e.g., [118,119]). Varied theories and hypotheses have been proposed about this sort of event (e.g., [120]). Similar accounts were considered for the phenomenon that occurred at the end of the Pleistocene worldwide, and in particular for southern Patagonia [121]. One that deserves to be mentioned is the “overkill” hypothesis, where prehistoric hunters played an important role (e.g., [122,123]), among others, a widely discussed approach cf. [124,125]. In the southern cone of South America, this alternative is considered to be one of its causes [121,126]. However, regional archaeologists agree that the PA population numbers were apparently low [127,128], and reliable archaeological sites with evidence on its exploitation are scarce. Besides, it is important to bear in mind that hunter-gatherer societies had very simple traditional techniques for hunting and surviving, and the number of projectiles found there is extremely low. Hence, in the complex environmental situation in southern Patagonia, even if there was some human influence on the mega mammals’ extinction, it would have been negligible.
When discussing the above-mentioned topics, it is important to bear in mind that close by the CM, paleomagnetic research performed on a terminal Pleistocene-early/middle Holocene sedimentary section in the Mylodon cave recorded a possible geomagnetic field (GMF) excursion [129]. It is defined as a deviation of the virtual geomagnetic pole by more than 40–45° from the geographic pole (Figure 18) and as causing a major deviation in the GMF’s behavior during a brief time span [130,131]. Additional evidence has been observed at the CM, and at other sites in Patagonia and southern South America [25,132,133,134,135], as well in other parts of the world see [136,137]. As previously discussed [138], paleomagnetic records with anomalous directions may be due to several reasons, but they may also reflect true GMF excursions. This kind of event might have caused major planetary environmental alterations [134,139,140]. In particular, the mechanisms triggered by these magnetic excursions might account for evolutionary discontinuities in humans and other animals, considering the increase in cosmic and ultraviolet radiation. Climate change and biomagnetic dysfunctions have also been proposed as mechanisms that may have accompanied the excursions and reversals [141]. As pointed out by Vizán and colleagues [142], during this sort of phenomenon, there are strong variations in the dipole moment [143] that might lead to a significant modulation of the magnetosphere and its shielding mechanism [144], which could be linked to a flux in the galactic cosmic rays reaching the Earth’s atmosphere, thus affecting the terrestrial weather system (e.g., [145]). Of these, increased ultraviolet radiation and biomagnetic disturbance are the most likely causes of biological modification. In this regard, from the perspective of human evolution, by performing a statistical test during the period 125.0–10.0 kya, when the Blake and Mungo excursions took place, Kopper and Papamarinopoulos [141] observed a strong correlation between both these events and times of abrupt physical and cultural changes in Old World prehistory [146,147,148]. The former coincides with the disappearance of Lower Paleolithic technology and the appearance of Middle Paleolithic (Mousterian) tools and Neanderthal Man, the latter synchronizes with the Middle/Upper Paleolithic boundary and the decline of Homo neanderthalensis, also in concurrence with the rapid high-amplitude climatic fluctuations that led to the Last Glacial Maximum [148]. By considering this and bearing in mind that during the last millennium of the Pleistocene and early Holocene, a significant number of records with intermediate and reverse directions around the world have been reported in North and South America, Eurasia, and Africa (e.g., [136,137,149,150]), a fact that led to the hypothesis of the possible global existence of this GMF anomaly during the time of the PA occupation. Hence, if the “Mylodon excursion” was a real fact, it is important to bear in mind that it happened during a time of crucial local and global transformations that gave rise to the Holocene [151], thus influencing the major paleoecological changes that occurred at that time. Like the reversals, it might also have had some effect on faunal extinctions [152,153,154]. Thus, the possible ~12–10.3 cal. kya short duration excursion cf. [152,155] might have had some kind of influence on the local archaeological process. In addition to this possible major cause of alterations, north of Ultima Esperanza in southern Chile, multidisciplinary research performed on sediments has suggested that a cosmic-impact generated biomass burning, climate change, and megafaunal extinctions at 12.8 cal. kya [156].
At this point, it would be valid to consider the survival or extinction of PA populations in the region, and lithic technology in particular may help explain this topic. In fact, because stone tools were made by craftsmen and were the results of a particular traditional technological knowledge (TTK) shared by a certain socio-cultural system, they constitute a “fossilized” record of human behavior [157,158]. Maintained in its members’ procedural memory [159,160], among other components, the technical expertise has its own recipes for fabrication, which involve the know-how for making them [157,161,162]. In summary, as part of the recipes for the fabrication of lithic tool making, shaped implements made by a determined socio-cultural group shared particular technological and morphological features. In the light of this theory, it is important to bear in mind that from a morphological and technological viewpoint, there are significant differences between the implements made by the FP makers and the lithic assemblages that came later in time (e.g., [3,4,5,92,115,116,163]). These dissimilarities are probably evidence that the PA technology was replaced by a different TTK shared by other socio-cultural groups. In fact, FP lithic assemblages were supplanted by others with triangular projectile points made using a different conception of bifacial reduction cf. [163,164] and the use of other forms of unifacial tools (cf. Figures 36–41) of [90], and Figures 8.24–8.29 of [95], some of which were made by employing a particular form of unifacial shaping [89,92]. This change allows us to hypothesize the extinction of PA technological knowledge and, consequently, the disappearance of the human groups that used it. This shift in technological patterns is likely the result of either socio-cultural and/or environmental changes between the Terminal Pleistocene and Early Holocene. Remarkable are the conclusions of a genetic study on South American human bone remains that link the earliest skeletons with those found at the Anzik burial and Clovis cache site [165]. Later, there was a population replacement at ~10 cal. kya with further continuity [166]. The oldest dates of the stemless triangular points in Patagonia are of a similar age [92,115,116] and, due to their distribution in the southern cone of South America, they probably represent human groups that came from the Andean region, probably Peru [53].
Finally, during the last decades several archaeological sites demonstrated the existence of pre-Clovis and pre-Fell occupations dated at ~15 cal. kya in North and South America (e.g., [167,168,169,170,171]) among others. Besides, by ~13–11.5 cal. kya the archaeological remains from that time show that a broad diversity of projectile points was employed in the Americas. Remarkably is that Clovis—the oldest fluted point—had spread throughout North America by ~13.0–12.7 cal. kya [172]. Similarly, in South America, the FP was a widespread lithic marker dating from between ~13–12 cal. kya [63,64,65,66,67,173]. Remarkably, the oldest dates of CM partially overlap the Clovis ages. Like classic Clovis points from the Great Plains, several FPs from Central and South America display true flutes, a reason why some scholars suggested its origin in the North American point [174]. However, most South America researchers conclude that they are technologically distinct in both morphology and reduction sequence [162,175,176,177]. Additionally, the early South American sites reveal marked diversity in lithic technology and subsistence pursuits [178,179]. The growing information on the New World human colonization shows that this process involved a complex scenario [180,181]. Actually, most scholars now believe that the Americas were peopled more than once [180], and that the different colonizing events produced remarkable technological and adaptive diversity, with the Pacific coast as the main initial entry route [169,182]. However, the Atlantic seaboard and the still-remaining Late Glacial Maximum exposed continental shelves might also have played a significant role in the expansion of the FP users [183,184,185,186]. Due to the similarities observed between the Central and South American FPs and the fishtailed specimens from eastern, and mainly southeastern North America, it is possible to hypothesize that there was a certain techno-morphological continuity in both areas [184,185,186]. Hence, as part of the various colonization events, hunter-gatherers using “fishtailed” points in North America might have been colonizing South America during the terminal Pleistocene by passing along the Central American isthmus and across the exposed continental shelves that existed in the Gulf of Mexico and the Caribbean Sea. The spreading from northern South America might have been followed two main “routes”: one through the Pacific slope and the Andean Cordillera in the west [186], and the other through Atlantic slope in the east [186,187]. In this scenario, despite it is located where the continent became narrower, and both “routes” may converge, in the colonization process of Patagonia, the FPs makers living at CM might belong to the major Pacific/Andean “route”, and being one of the southernmost sites located in western South America witnessed by the finds of Salar Punta Negra [188], Quebrada Santa Julia [189], Valiente [190], Tagua-Tagua [191], and Salto Chico in Chile [167], and those FPs surficial finds performed in the Salta, Catamarca, Mendoza, Neuquén, and Río Negro provinces in Argentina (e.g., [192,193,194,195]).

5. Final Remarks

In summary, the re-discovery and excavations at the CM became a turning point in the study of the oldest hunter-gatherer societies living in southern Patagonia, as well in the development and continuity of the regional research and scientific community. Due to the looting and excavations that occurred at the end of the 19th century, a large portion of the CM’s archaeologically excavatable surface was destroyed. However, intact deposits yielded evidence that is still useful for reliably discovering the regional and continental human past regarding the peopling of the Americas. Finds from there have been extremely useful for discussing diverse paleo-ecological and archaeological topics and have extended the knowledge and discussions about different Pleistocene scientific topics, mainly to understand the earliest regional human occupations. The finds performed at CM provided insights into the paleoenvironment prevailing at the time of the PA foragers that interacted with an important Late Pleistocene faunal diversity, some of which were exploited for consumption. In this regard, with highly controlled chronological data, the sealed and direct association between the bones of extant and extinct fauna allowed us to discover and discuss the consumption habits of the early foragers who used FPs as part of their weaponry. In particular, from the stone tool analysis perspective, it is one of the few sites yielding FPs in association with the assemblages accompanying them. By this way, the aforementioned investigations expanded our understanding and knowledge of the earliest colonizers of southern Patagonia and human-megafauna interaction before extinction.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2571-550X/2/3/28/s1, Figure S1: Pollen and spore diagram from the samples taken at CM. The numbers in parentheses on level three indicate depths in cm below the datum. Modified after Figure 4 of [26]. Figure S2: Comparative graphs of the daily temperature recorded at the CM during the morning (a), at midday (b), and during the afternoon (c). References—C: Cloudy, PC: Partly cloudy, R: Rainy, S: Sunny.

Funding

The first expedition was carried out using small donations from a local bus company (round trip ticket PA-PN-PA) and grocery stores from Punta Arenas, as well as HGN’s own personal funds. Furthermore, due to the potential magnitude of the findings [13], the IP and HGN, respectively, acquired small grants from the Bird Foundation (New York, NY, USA) and Sigma Xi, the Scientific Research Society, in 1988 and 1989. In 1992, HGN was granted a three-year project by the Wenner-Gren Foundation for Anthropological Research. In addition, other institutions allowed us to receive different kinds of support, mainly related to the field-work and radiocarbon data. They are as follows: CONICET (Argentina), the Paleoindian Program (Smithsonian Institution, Washington, DC, USA), and the Mongoloid Dispersal Project (Japan). The sixth expedition was financed by a grant given by the National Geographic Society to the IP. It is worth recalling that during this project, the significance of the CM was used to carry out a broader undertaking in order to discover the CB’s paleontology and archaeology along with the participation of a number of national and international scholars.

Acknowledgments

My special thanks to: J. Llorca, former editor of Quaternary for his invitation to contribute in this volume; E. Montoya, B. Whitney, and V. Rull for the kindness of allow to participate in this special issue; many Patagonian friends and colleagues who were so helpful and generous during the nice times that we spend during the Cueva del Medio field works; the institutions that supported logistic, and financial support for this research; an anonymous reviewer, and especially Michael J. O’Brien provided useful observations, help and cooperation during the edition of this paper.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Graf, K.E.; Ketron, C.V.; Waters, M.R. Paleoamerican Odissey; Center for the Study of the first Americans: College Station, TX, USA, 2013. [Google Scholar]
  2. Nami, H.G. Arqueología del último milenio del Pleistoceno en el Cono Sur de Sudamérica, puntas de proyectil y observaciones sobre tecnología Paleoindia en el Nuevo Mundo. In Peuplement et Modalités D’occupation de L’amérique du Sud: L’apport de la Technologie Lithique; Farias, M., Lourdeau, A., Eds.; @rchéo-éditions.com: Prigonrieux, France, 2014; pp. 279–336. [Google Scholar]
  3. Bird, J. Before Magellan. Nat. Hist. 1938, 41, 16–28. [Google Scholar]
  4. Bird, J. The archaeology of Patagonia. In Handbook of South American Indians; Steward, J., Ed.; Smithsonian Institutio: Washington, DC, USA, 1946; pp. 17–24. [Google Scholar]
  5. Bird, J. Travels and Archaeology in South Chile; Hyslop, J., Ed.; Iowa University Press: Des Moines, IA, USA, 1988. [Google Scholar]
  6. Sanguinetti, A.C.; Borrero, L.A. Niveles con Fauna Extinta de la Cueva de Las Buitreras (Santa Cruz). Relac. Soc. Argent. Antropol. 1977, 11, 167–175. [Google Scholar]
  7. Massone, M. Los cazadores paleoindios de Tres Arroyos (Tierra del Fuego). An. Inst. Patagon. 1987, 17, 47–60. [Google Scholar]
  8. Nami, H.G. Cueva del Medio: Perspectivas Arqueológicas para la Patagonia Austral. An. Inst. Patagon. 1987, 17, 71–106. [Google Scholar]
  9. Prieto, A. Cazadores tempranos y tardíos en la Cueva del Lago Sofia. An. Inst. Patagon. 1991, 20, 75–99. [Google Scholar]
  10. Borrero, L.A. Cazadores de Mylodon en Patagonia austral. In New Evidences for the Pleistocene Peopling of the Americas; Bryan, A.L., Ed.; Center for the Study of Early Man: Orono, MI, USA, 1986; pp. 281–294. [Google Scholar]
  11. Borrero, L.A.; Lanata, J.L.; Borella, F. Reestudiando huesos: Nuevas consideraciones sobre sitios de Ultima Esperanza. An. Inst. Patagon. 1988, 18, 133–156. [Google Scholar]
  12. Orquera, L.A. Advances in the archaeology of the Pampa and Patagonia. J. World Prehist. 1987, 1, 333–413. [Google Scholar] [CrossRef]
  13. Nami, H.G. Excavación arqueológica y hallazgo de una punta de proyectil “Fell I” en la Cueva del Medio, Seno de Ultima Esperanza, Chile. Informe preliminar. An. Inst. Patagon. 1985, 16, 103–109. [Google Scholar]
  14. Nami, H.G. Cueva del Medio: A Significant Paleoindian Site in Southern South America. Curr. Res. Pleistocene 1987, 4, 157–158. [Google Scholar]
  15. Nami, H.G. Nuevos Antecedentes sobre Investigaciones Arqueológicas en el Area de Cerro Benítez (Ultima Esperanza, Magallanes). An. Inst. Patagon. 1990, 19, 125–132. [Google Scholar]
  16. Nami, H.G. Resumen de las actividades y nuevos datos obtenidos en la quinta campana de investigaciones arqueologicas en Ultima Esperanza, Chile. Palimpsesto 1992, 2, 123–132. [Google Scholar]
  17. Nami, H.G. Las excavaciones arqueológicas y los hallazgos de fauna extinta en el Seno de Ultima Esperanza, Chile Explotación de Recursos Faunísticos en Sistemas Adaptativos Americanos. Arq. Cont. 1993, 4, 123–133. [Google Scholar]
  18. Nami, H.G. Reseña sobre los avances de la arqueología finipleistocénica del extremo sur. Chungará 1994, 26, 145–163. [Google Scholar]
  19. Nami, H.G. New Assessments of Early Human Occupations in the Southern Cone. In Prehistoric Mongoloid Dispersals; Akazawa, T., Szathmáry, E.J.E., Eds.; Oxford University Press: Oxford, UK, 1996; pp. 254–269. [Google Scholar]
  20. Menegaz, A.; Nami, H.G. Análisis de los materiales faunísticos del sitio arqueológico “Cueva del Medio”, Seno de Ultima Esperanza, Chile. In VI Jornadas Argentinas de Paleontología de Vertebrados; Instituto y Museo de Ciencias Naturales; Universidad Nacional de San Juan: San Juan, Argentina, 1989; pp. 80–81. [Google Scholar]
  21. Menegaz, A.; Nami, H.G. Late Pleistocene Faunal Diversity in Ultima Esperanza (Chile). Further Data from Cueva del Medio. Curr. Res. Pleistocene 1994, 11, 93–94. [Google Scholar]
  22. Nami, H.G.; Menegaz, A. Cueva del Medio: Aportes para el conocimiento de la diversidad faunística hacia el Pleistoceno Holoceno en Patagonia Austral. An. Inst. Patagon. 1991, 21, 117–132. [Google Scholar]
  23. Nami, H.G. Excavaciones y Hallazgos de Fauna Extinguida en Cueva Del Medio, Ultima Esperanza, Chile. Presented at II Jornadas de Arqueología de la Patagonia, Puerto Madryn, Argentina, 1996. [Google Scholar]
  24. Nami, H.G.; Nakamura, T. Cronologia radiocarbonica con AMS sobre muestras de huesos procedentes del sitio Cueva del Medio (Ultima Esperanza, Chile). An. Inst. Patagon. 1995, 23, 125–133. [Google Scholar]
  25. Nami, H.G.; Sinito, A.M.; Gogorza, C. Primeros resultados de los estudios paleomagneticos en sedimentos de Cueva del Medio (Ultima Esperanza, Chile). An. Inst. Patagon. 1995, 23, 135–142. [Google Scholar]
  26. Nami, H.G.; Heusser, C.J. Cueva del Medio: A Paleoindian Site and Its Environmental Setting in Southern South America. Archaeol. Discov. 2015, 3, 62–71. [Google Scholar] [CrossRef] [Green Version]
  27. Alberdi, M.T.; Prieto, A. Hippidion (Mammalia, Perissodactyla) de las cuevas de las provincias de Magallanes y Tierra del Fuego. An. Inst. Patagon. 2000, 28, 147–171. [Google Scholar]
  28. Martin, F.M. Tafonomía de la Transición Pleistoceno-Holoceno en Fuego-Patagonia. In Interacción Entre Humanos y Carnívoros y su Importancia Como Agentes en la Formación del Registro Fósil; Ediciones de la Universidad de Magallanes: Punta Arenas, Chile, 2013. [Google Scholar]
  29. Martin, F.M.; Todisco, D.; Rodet, J.; San Román, M.; Morello, F.; Prevosti, F.; Stern, C.; Borrero, L.A. Nuevas excavaciones en Cueva del Medio. Procesos de formación de la cueva y avances en los estudios de interacción entre cazadores recolectores y fauna extinta (Pleistoceno final, Patagonia Meridional). Magallania 2015, 43, 165–189. [Google Scholar]
  30. Vásquez, C.A.; Nami, H.G.; Rapalini, A.E. Magnetic Sourcing of Obsidians in Southern South America. J. Archaeol. Sci. 2001, 28, 613–618. [Google Scholar] [CrossRef]
  31. Labarca Encina, R. La Subsistencia de Los Cazadores Recolectores de Patagonia Meridional Chilena Durante la Transición Pleistoceno-Holoceno: Un Enfoque Integrador Desde la Zooarqueología. Ph.D. Thesis, Facultad de Ciencias Sociales, Universidad Nacional del Centro, Olavarria, Argentina, 2016. [Google Scholar]
  32. Mouginot, J.; Rignot, E. Ice motion of the Patagonian Icefields of South America: 1984–2014. Geophys. Res. Lett. 2015, 42, 1441–1449. [Google Scholar] [CrossRef]
  33. Ramos, V.A.; Duncan Keppie, J. Laurentia-Gondwana Connections before Pangea; Geological Society of America: Washington, DC, USA, 1999. [Google Scholar]
  34. Sutcliffe, A.J. On the Track of Ice Age Mammals; Harvard University Press: Cambridge, UK, 1986. [Google Scholar]
  35. Martinic, M. La Cueva del Milodon (Ultima Esperanza, Patagonia chilena). Un siglo de descubrimientos y estudios referidos a la vida primitiva en el sur de America. J. Soc. Am. 1996, 82, 311–323. [Google Scholar]
  36. Emperaire, J.; Laming, A. La grotte du Mylodon (Patagonie occidentale). J. Soc. Am. 1954, 43, 173–206. [Google Scholar] [CrossRef]
  37. Ameghino, F. Première notice sur le Neomylodon Listai, un représentant vivant des anciens Édentés Gravigrades fossiles de l’Argentine. Nat. Sci. 1898, 81, 324. [Google Scholar]
  38. Hauthal, R.; Roth, S.; Lehmann-Nitsche, R. El Mamifero Misterioso de la Patagonia “Grypotherium domesticum”. Rev. Mus. Plata 1899, 9, 409–420. [Google Scholar]
  39. Waters, M.R. Principles of Geoarchaeology: A North American Perspective; University of Arizona Press: Tucson, AZ, USA, 1992. [Google Scholar]
  40. Ones, B. Cave sediments. In Encyclopedia of Sediments and Sedimentary Rocks; Middleton, G.V., Church, M.J., Coniglio, M., Hardie, L.A., Longstaffe, F.J., Eds.; Springer: Dordrecht, The Netherlands, 2003; pp. 172–173. [Google Scholar]
  41. Dott, R.H.J. Episodic sedimentation-How normal is average? How rare is rare? Does it matter? J. Sediment Petrol. 1983, 53, 5–23. [Google Scholar] [CrossRef]
  42. Feathers, J.; Nami, H.G. Luminescence Dating of Late Pleistocene and Holocene Sediments in Uruguay. Lat. Am. Antiq. 2018, 29, 495–513. [Google Scholar] [CrossRef]
  43. Wheat, J.B. The Olsen-Chubbuck Site: A Paleo-Indian Bison Kill Memoirs 26; Society for American Archaeology: Washington, DC, USA, 1972. [Google Scholar]
  44. Frison, G.C.; Stanford, D.J. The Agate Basin Site; Academic Press: New York, NY, USA, 1982. [Google Scholar]
  45. Politis, G.G.; Messineo, P.G.; Stafford Jr, T.W.; Lindsey, E.L. Campo Laborde: A Late Pleistocene Giant Ground Sloth Kill and Butchering Site in the Pampas. Sci. Adv. 2019, 5, 4546. [Google Scholar] [CrossRef]
  46. Binford, L.R. Bones: Ancient Men and Modern Myths; Academic Press: New York, NY, USA, 1981. [Google Scholar]
  47. Lyman, R.L. Vertebrate Taphonomy; Cambridge University Press: Cambridge, UK, 1994. [Google Scholar]
  48. Menegaz, A.N.; Nami, H.G.; Senatore, M.X. Alteraciones de los restos faunísticos óseos de Cueva del Medio: Un análisis preliminar, Actas y Memorias del XI Congreso Nacional de Arqueología Argentina. Resúm. Rev. Mus. Hist. Nat. San Rafael 1994, 13, 35–36. [Google Scholar]
  49. López Mendoza, P.; Rojas Mondaca, O. El Pleistoceno de la Cuenca de Calama. Evidencias Paleontológicas y Arqueológicas. In Monografías del Museo de Historia Natural y Cultural del Desierto de Atacama; Gráfica LOM: Santiago, Chile, 2018. [Google Scholar]
  50. Martin, F.; Borrero, L.A. Climate change, availability of territory, and Late Pleistocene human exploration of Ultima Esperanza, South Chile. Quat. Int. 2017, 428, 86–95. [Google Scholar] [CrossRef]
  51. Lynch, T. Glacial-Age Man in South America? A Critical Review. Am. Antiq. 1990, 55, 12–36. [Google Scholar] [CrossRef]
  52. Borrero, L.A. Comment to “Caballos Fósiles”. Int. Antiq. 2005, 6, 231–233. [Google Scholar]
  53. Salemme, M.; Miotti, L. Archaeological Hunter-Gatherer Landscapes Since the Latest Pleistocene in Fuego-Patagonia. In The Late Cenozoic of Patagonia and Tierra del Fuego; Rabassa, J., Ed.; Elsevier: New York, NY, USA, 2008; pp. 437–483. [Google Scholar]
  54. Massone, M. Los cazadores después del hielo. In Santiago de Chile: Dirección de Bibliotecas, Archivo y Museos; Centro de Investigaciones Diego Barros Arana: Santiago, Chile, 2004. [Google Scholar]
  55. Massone, M.; Prieto, A. Evaluación de la modalidad cultural Fell 1 en Magallanes. Chungara 2004, 36, 303–315. [Google Scholar] [CrossRef]
  56. Messineo, P. Technological Organization in Hunting/Scavenging and Butchering Sites of Megamammals in the Pampa Grasslands (Argentina). Paleo Am. 2015, 1, 343–359. [Google Scholar] [CrossRef]
  57. Paunero, R.; Castro, A. Análisis lítico y funcionalidad del componente inferior del sitio Cueva 1. Localidad arqueológica Cerro Tres Tetas. Provincia de Santa Cruz. Argentina. An. Inst. Patagon. 2001, 29, 189–206. [Google Scholar]
  58. Politis, G.G.; Prates, L.; Pérez, S.I. El Poblamiento de America. Arqueología y Bio-Antropologia de los Primeros Americanos; Eudeba: Buenos Aires, Argentina, 2009. [Google Scholar]
  59. Martínez, G.; Gutiérrez, M.; Messineo, P.; Kaufmann, C.; Rafuse, D.J. Subsistence strategies in Argentina during the Late Pleistocene and Early Holocene. Quat. Sci. Rev. 2016, 144, 51–65. [Google Scholar] [CrossRef]
  60. Nami, H.G. Late Pleistocene Technology in the New World: Bone Artifacts from Cueva del Medio and other Sites in the Southern Cone of South America. In Ancient and Modern Bone Artefacts from America to Russia Cultural, Technological and Functional Signature; Legrand-Pineau, A., Sidéra, I., Buc, N., David, E., Scheinsohn, V., Eds.; BAR International Series: Oxford, UK, 2010; pp. 279–286. [Google Scholar]
  61. Metcalf, J.L.; Turney, C.; Barnett, R.; Martin, F.; Bray, S.C.; Vilstrup, J.T.; Orlando, L.; Salas-Gismondi, R.; Loponte, D.; Medina, M.; et al. Synergistic roles of climate warming and human occupation in Patagonian megafaunal extinctions during the Last Deglaciation. Sci. Adv. 2016, 2, 1–8. [Google Scholar] [CrossRef]
  62. Reimer, P.J.; Bard, E.; Bayliss, A.; Beck, J.W.; Blackwell, P.G.; Ramsey, C.B.; Buck, C.E.; Cheng, H.; Edwards, R.L.; Friedrich, M.; et al. Intcal13 and Marine13 Radiocarbon Age Calibration Curves 0–50,000 Years Cal Bp. Radiocarbon 2013, 55, 1869–1887. [Google Scholar] [CrossRef]
  63. Nami, H.G. Research in the Middle Negro River Basin (Uruguay) and the Paleoindian Occupation of the Southern Cone. Curr. Anthropol. 2007, 48, 164–176. [Google Scholar] [CrossRef]
  64. Nami, H.G. Silcrete as a Valuable Resource for Stone Tool Manufacture and Its Use by Paleo-American Hunter-Gatherers in Southeastern South America. J. Archaeol. Sci. Rep. 2016, 15, 539–560. [Google Scholar] [CrossRef]
  65. Nami, H.G.; Stanford, D.J. Dating the Peopling of Northwestern South America: An AMS Date from El Inga Site, Highland Ecuador. Paleo Am. 2016, 2, 60–63. [Google Scholar] [CrossRef]
  66. Yataco Capcha, J.; Nami, H.G. A Re-Evaluation of Paleo-American Artifacts from Jaywamachay Rockshelter, Ayacucho Valley, Peru. Paleo Am. 2016, 2, 368–372. [Google Scholar]
  67. Waters, M.R.; Amorosi, T.; Stafford, T. Redating Fell’s Cave, Chile and the chronological placement of the Fishtail projectile point. Am. Antiq. 2015, 80, 80–376. [Google Scholar] [CrossRef]
  68. Aschero, C.A. Ensayo Para Una Clasificación Morfológica de Artefactos Líticos Aplicada a Estudios Tipológicos Comparativos; Science Open: Burlington, MA, USA, 1975. [Google Scholar]
  69. Nami, H.G.; Casé, A.M. The Raw Material used by the Paleoindian of the Cueva del Medio. Curr. Res. Pleistocene 1988, 5, 31–32. [Google Scholar]
  70. Nami, H.G. Secuencias de Reducción Bifaciales Paleoindias y Puntas Fell en el Valle del Ilaló (Ecuador): Observaciones para comprender la tecnologia lítica pleistocénica en Sudamérica. In Peuplement et Modalités D’occupation de L’amérique du Sud: L’apport de la Technologie Lithique; Farias, M., Lourdeau, A., Eds.; @rchéo-éditions.com: Prigonrieux, France, 2014; pp. 179–220. [Google Scholar]
  71. Nami, H.G. Paleo American Artifacts from Cerro Largo, Northeastern Uruguay. Paleo Am. 2015, 1, 288–292. [Google Scholar]
  72. Borrero, L.A.; Nami, H.G. Arqueología en Piedra del Aguila: Prospecciones y propuestas metodológicas. II) Piedra del Aguila: Análisis de los materiales de superficie. Præhistoria 1996, 2, 19–34. [Google Scholar]
  73. Crivelli Montero, E.A.; Fernández, M.M. Paleoindian bedrock engravings at Epullán Grande Cave (northern Patagonia, Argentina). Rock Art Res. 1996, 13, 124–128. [Google Scholar]
  74. Crivelli Montero, E.A.; Pardiñas, U.F.J.; Fernández, M.M.; Bogazzi, M.; Chauvin, A.; Fernández, V.M.; Lezcano, M. La Cueva Epullán Grande (provincia del Neuquén, Argentina) Informe de avance. Præhistoria 1996, 2, 185–265. [Google Scholar]
  75. Bird, J. A comparison of south Chilean and Ecuatorial “fishtail” projectile points. Kroeber Antiq. Soc. Pap. 1969, 40, 52–71. [Google Scholar]
  76. Weitzel, C.; Flegenheimer, N.; Martínez, J.; Colombo, M. Breakage Patterns on Fishtail Projectile Points. Ethnoarchaeology 2014, 6, 81–102. [Google Scholar]
  77. Nami, H.G. Experimentos para explorar la secuencia de reducción Fell de la Patagonia Austral. Magallania 2003, 30, 107–138. [Google Scholar]
  78. Nami, H.G. Tecnología Paleoindia de Sudamérica: Nuevos experimentos y observaciones para conocer la secuencia de reducción Fell. Orígenes 2010, 9, 1–40. [Google Scholar]
  79. Nami, H.G. Investigaciones actualístico-experimentales para aproximarse a la tecnología paleoindia: Comparación de las secuencias de reducción Folsom-Lindenmeier y Fell de la Patagonia. In La Investigación Experimental Aplicada a la Arqueología; Morgado, A., Preysler, J.B., González, D.G., Eds.; Universidad de Granada-Universidad Autónoma de Madrid: Ronda, Spain, 2011; pp. 97–103. [Google Scholar]
  80. Orquera, L.A. Discussion. In Primeras Jornadas de Tipología y Tecnología Líticas; Centro de Investigaciones Antroológicas: Buenos Aires, Argentina, 1980; pp. 34–35. [Google Scholar]
  81. Jodry, M.A. Folsom Technological and Socioeconomic Strategies: Views from Stewart’s Cattle Guard and the Upper Rio Grande Basin, Colorado. Ph.D. Thesis, American University, Washington, DC, USA, 1999. [Google Scholar]
  82. Nami, H.G.; Borella, F. Investigaciones actualísticas-experimentales aplicadas a la interpretación de huellas de cercenamiento en restos arqueofaunísticos de cetáceos de Tierra del Fuego. An. Inst. Patagon. 1999, 27, 239–253. [Google Scholar]
  83. Dibble, H.L. Middle paleolithic scraper reduction: Background, clarification, and review of the evidence to date. J. Archaeol. Method Theory 1995, 2, 299–368. [Google Scholar] [CrossRef]
  84. Baena Preysler, J.; Carrión Santafé, E. Experimental Approach to the Function and Technology of Quina Side-Scrapers. In Experiment and Interpretation of Traditional Technologies: Essays in Honor of Errett Callahan; Nami, H.G., Ed.; Scientific Research Publishing: Irvine, CA, USA, 2010; pp. 171–202. [Google Scholar]
  85. Binford, L.R. Organization and Formation Processes: Looking at Curated Technologies. J. Anthropol. Res. 1979, 35, 255–273. [Google Scholar] [CrossRef]
  86. Nash, H. Is Curation a Useful Heuristic. In Stone Tools: Theoretical Insights Into Human Prehistory; Odell, G.H., Ed.; Plenum Press: New York, NY, USA, 1996; pp. 81–99. [Google Scholar]
  87. Nami, H.G. Archaelogy, Paleoindian Research and Lithic Technology in the Middle Negro River, Central Uruguay. Archaeol. Discov. 2013, 1, 1–22. [Google Scholar] [CrossRef] [Green Version]
  88. Nami, H.G. Experimental Observations on some Non-Optimal Materials from Southern South America. Lithic Technol. 2015, 40, 128–146. [Google Scholar] [CrossRef]
  89. Civalero, M.T.; Nami, H.G. Experimentos y Esquemas Diacríticos para Explorar Técnicas de Talla Unifacial del Holoceno Temprano en el NO de Santa Cruz. Rev. Mus. Antiq. 2019. under review. [Google Scholar]
  90. Mansur-Franchomme, M.E. Traces D’utilisation et Technologie Lithique: Exemples de la Patagonie. Ph.D. Thesis, Bordeaux I University, Bordeaux, France, 1983. [Google Scholar]
  91. Cardich, A. Arqueología de Los Toldos y El Ceibo (Provincia de Santa Cruz, Argentina). Estud. Atac. 1987, 8, 98–117. [Google Scholar] [CrossRef]
  92. Nami, H.G.; Civalero, M.T. Distinctive Unifacial Technology during the Early Holocene in Southern South America. Archaeol. Discov. 2017, 5, 101–115. [Google Scholar] [CrossRef] [Green Version]
  93. Hermo, D.; Mosquera, B.; Vargas Gariglio, J.; Pérez, A. Indicadores tecnológicos de colonización en contextos de superficie del Macizo del Deseado (Santa Cruz, Argentina). In Libro de Resúmenes del 9 Simposio Internacional el Hombre Temprano en América La Gente y Sus Lugares; CONICET Documental: Buenos Aires, Argentina, 2018; pp. 66–67. [Google Scholar]
  94. García, A. Los Primeros Pobladores de los Andes Centrales Argentinos: Una Mirada a los Estudios Sobre los Grupos Cazadores-Recolectores Tempranos de San Juan y Mendoza; Zeta Editores: Mendoza, Argentina, 2004. [Google Scholar]
  95. Cattáneo, R. Tecnología Lítica del Pleistoceno Final /Holoceno Medio Un Estudio de los Cazadores-Recolectores de la Patagonia Austral (Argentina); BAR International Series: Oxford, UK, 2006. [Google Scholar]
  96. Paunero, R. The Cerro Tres Tetas (C3T) Locality in in the Central Plat al Plat al Plateau of Santa Cruz, Argentina. In Where the South Winds Blow: Ancient Evidence of Paleo South Americans; Miotti, L., Salemme, M., Flegenheimer, N., Eds.; Center for the Study of the first Americans: College Station, TX, USA, 2003; pp. 133–140. [Google Scholar]
  97. Paunero, R. La colonización humana de la meseta central de Santa Cruz durante el Pleistoceno final: Indicadores arqueológicos, referentes estratigráficos y nuevas evidencias. In Arqueología de Patagonia: Una Mirada Desde el Último Confín; Álvarez, M., Piana, E., Vázquez, M., Mansur, E., Salemme, M., Santiago, F., Eds.; Utopías: Ushuaia, Argentina, 2009; pp. 85–100. [Google Scholar]
  98. Nami, H.G.; Florines, A.; Toscano, A. New Paleoindian Finds, Further Fell Points Data, and Technological Observations from Uruguay: Implications for the Human Peopling in Southeastern South America. Archaeol. Discov. 2018, 6, 21–37. [Google Scholar] [CrossRef] [Green Version]
  99. Schmitz, P.I. Opovoamento pleistocênico do Brasil. Rev. Arq. Am. 1990, 1, 37–60. [Google Scholar]
  100. Araujo, A.G.M. On Vastness and Variability: Cultural Transmission, Historicity, and the Paleoindian Record in Eastern South America. An. Acad. Bras. Ciênc. 2015, 87, 1239–1258. [Google Scholar] [CrossRef]
  101. Loponte, D.; Carbonera, M.; Silvestre, R. Fishtail Projectile Points from South America: The Brazilian Record. Archaeol. Discov. 2015, 3, 85–103. [Google Scholar] [CrossRef] [Green Version]
  102. Loponte, D.; Okumura, M.; Carbonera, M. New Records of Fishtail Projectile Points from Brazil and Its Implications for Its Peopling. J. Lithic Stud. 2016, 3, 63–85. [Google Scholar] [CrossRef]
  103. Nami, H.G. Exceptional Fell Projectile Points from Uruguay: More Data on Paleoindian Technology in the Southern Cone. Curr. Res. Pleistocene 2011, 28, 112–116. [Google Scholar]
  104. Barbosa, A.; Universidade Católica de Goiás, Brasília, Brazil. Personal communication, 1995.
  105. Meggers, B. Department of Anthropology, National Museum of Natural History; Smithsonian Institution: Washington, DC, USA, 1995. [Google Scholar]
  106. Adams, J.M. Map of the Vegetation in South America during 10 Ka Years BP. Unpublished. 1996. [Google Scholar]
  107. Adams, J.M.; Faure, H. Preliminary Vegetation Maps of the World since the Last Glacial Maximum: An Aid to Archaeological Understanding. J. Archaeol. Sci. 1997, 24, 623–647. [Google Scholar] [CrossRef]
  108. McCulloch, R.D.; Davies, S.J. Late-glacial and Holocene palaeoenvironmental change in the central Strait of Magellan, southern Patagonia. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2001, 173, 143–173. [Google Scholar] [CrossRef]
  109. Loponte, D.; Carbonera, M. Paleoamericans in Northeast Argentina. Archaeol. Discov. 2017, 5, 79–94. [Google Scholar] [CrossRef] [Green Version]
  110. Toggweiler, J.R. Shifting Westerlies. Science 2009, 323, 1434–1435. [Google Scholar] [CrossRef]
  111. Menviel, L.; Spence, P.; Yu, J.; Chamberlain, M.A.; Matear, R.J.; Meissner, K.J.; England, M.H. Southern Hemisphere westerlies as a driver of the early deglacial atmospheric CO2 rise. Nat. Commun. 2018, 9, 2503. [Google Scholar] [CrossRef]
  112. Nami, H.G.; Compagnucci, R.; Castañeda, E.; Cuadrado Woroszylo, M. El Paleoclima de los últimos 15,000 años a través de la información arqueológica. In IX Congreso Latinoamericano e Ibérico de Meteorología, VII Congreso Argentino de Meteorología; Centro Argentino de Meteorólogos: Buenos Aires, Argentina, 2001. [Google Scholar]
  113. Compagnucci, R.; Nami, H.G.; Castañeda, E.; Cuadrado Woroszylo, M. Climatic and human occupation changes since 15000 BP in southern Patagonia. In European Research Conference on “Achieving Climate Predictibility Using Paleoclimatic Data”; Castelvecchio: Pascoli, Italy, 2001. [Google Scholar]
  114. Meltzer, D.J. Issues of Scale, Demography, and Landscape Learning. In The Settlement of the American Continents: A Multidisciplinary Approach to Human Biogeography; Barton, C.M., Clark, G.A., Yesner, D.R., Pearson, G.A., Eds.; The University of Arizona Press: Tucson, AZ, USA, 2004; pp. 123–137. [Google Scholar]
  115. Aguerre, A.M. La cueva 4 de La Martita y las ocupaciones de 8000 años (Santa Cruz, Argentina). In Arqueología y Paleoambiente en la Patagonia Santacruceña Argentina; Aguerre, A.M., Ed.; Búsqueda de Ayllu: Buenos Aires, Argentina, 2003; pp. 27–62. [Google Scholar]
  116. Gradín, C.J.; Aguerre, A.M. Contribución a la Arqueología del Río Pinturas; Búsqueda de Ayllu: Concepción del Uruguay, Argentina, 1994. [Google Scholar]
  117. Nami, H.G. Archaeological Research in the Argentinean Río Chico Basin. Cur. Ant. 1995, 36, 661–664. [Google Scholar] [CrossRef]
  118. Diamond, J.M.; Ashmole, N.P.; Purves, P.E. The Present, Past and Future of Human-Caused Extinctions. Philos. Trans. R. Soc. Lond. B Biol. Sci. 1989, 1228, 469–477. [Google Scholar] [CrossRef]
  119. Wooldridge, S.A. Mass extinctions past and present: A unifying hypothesis. Biogeosci. Discuss. 2008, 5, 2401–2423. [Google Scholar] [CrossRef]
  120. Proenca, V.; Pereira, H.M. Comparing Extinction Rates: Past, Present, and Future. In Encyclopedia of Biodiversity; Levin, S.A., Ed.; Academic Press: Oxford, UK, 2013; Volume 2, pp. 167–176. [Google Scholar] [CrossRef]
  121. Villavicencio, N.A.; Lindsey, E.L.; Martin, F.M.; Borrero, L.A.; Moreno, P.I.; Marshall, C.R. Combination of humans, climate, and vegetation change triggered Late Quaternary megafauna extinction in the Última Esperanza region, southern Patagonia Chile. Ecography 2016, 39, 125–140. [Google Scholar] [CrossRef]
  122. Martin, P.S. Prehistoric overkill. In Prehistoric Extinctions: The Search for a Cause; Martin, P.S., Wright, H.E., Eds.; Yale University Press: New Haven, CT, USA, 1967; pp. 75–120. [Google Scholar]
  123. Martin, P.S. The Discovery of America. Science 1973, 179, 969–974. [Google Scholar] [CrossRef] [Green Version]
  124. Grayson, D.K.; Meltzer, D.J. A requiem for North American overkill. J. Archaeol. Sci. 2003, 30, 585–593. [Google Scholar] [CrossRef] [Green Version]
  125. Fiedel, S.; Haynes, H. A premature burial: Comments on Grayson and Meltzer’s “Requiem for overkill”. J. Archaeol. Sci. 2004, 31, 121–131. [Google Scholar] [CrossRef]
  126. Prado, J.L.; Alberdi, M.T. Fossil Horses of South America: Phylogeny, Systemics and Ecology; Springer: New York, NY, USA, 2017. [Google Scholar]
  127. Miotti, L.; Salemme, M. Poblamiento, movilidad y territorios entre las sociedades cazadoras-recolectoras de Patagonia. Complutum 2004, 15, 177–206. [Google Scholar]
  128. Scheinsohn, V. Hunter-Gatherer Archaeology in South America. Annu. Rev. Anthropol. 2003, 32, 339–361. [Google Scholar] [CrossRef]
  129. Nami, H.G. Holocene Geomagnetic Excursion at Mylodon Cave, Ultima Esperanza Chile. J. Geomagn. Geoelectr. 1995, 47, 1325–1332. [Google Scholar] [CrossRef]
  130. Twyman, R.M. Geomagnetic Excursions and Secular Variations. In Encyclopedia of Quaternary Science; Elias, S.A., Ed.; Elsevier: Oxford, UK, 2007; pp. 717–720. [Google Scholar]
  131. Laj, C.; Channel, J.E.T. Geomagnetic excursions. In Treatise on Geophysics; Kono, M., Ed.; Elsevier: Amsterdam, The Netherlands, 2007; pp. 373–416. [Google Scholar] [CrossRef]
  132. Nami, H.G. Nuevo registro de la excursión geomagnética post-pleistocénica en la cueva de Las Buitreras (Provincia de Santa Cruz). Præhistoria 1999, 3, 163–168. [Google Scholar]
  133. Nami, H.G. Possible Holocene Excursion of the Earth’s Magnetic Field in Southern South America: New Records from Archaeological Sites in Argentina. Earth Planets Space 1999, 51, 175–191. [Google Scholar] [CrossRef]
  134. Nami, H.G. New Detailed Holocene Paleomagnetic Records with Anomalous Geomagnetic Field Behavior in Argentina. Geoacta 2012, 37, 83–116. [Google Scholar]
  135. Nami, H.G. Paleomagnetic Results from Archaeological Sites in Argentinean Patagonia: Evidence for the Holocene Geomagnetic Excursions in Southern South America and Its Chronostratigraphic Implications. Archaeol. Discov. 2019, 7, 127–154. [Google Scholar] [CrossRef] [Green Version]
  136. Nami, H.G.; Vásquez, C.A.; Durán, V.A. Detailed Early Holocene (10.3 cal kybp) Paleomagnetic Record with Anomalous Directions from Mendoza Province, Western Argentina. Latinmag Lett. 2017, 7, 1–17. [Google Scholar]
  137. Nami, H.G.; Vásquez, C.A.; Wadley, L.; de la Peña, P. Detailed Paleomagnetic Record at Rose Cottage Cave, South Africa: Implications for the Holocene Geomagnetic Field Behavior and Chronostratigraphy. Archaeol. Discov. 2019. under revision. [Google Scholar]
  138. Nami, H.G.; de la Peña, P.; Vásquez, C.; Feathers, J.; Wurz, S. Paleomagnetic Results and New Dates from Late Pleistocene and Holocene Deposits from Klasies River Cave 1, South Africa. Afr. J. Sci. 2016, 112, 2016–2051. [Google Scholar] [CrossRef]
  139. Kuznetsova, N.D.; Kuznetsov, V.V. Implications of Volcanism and Geomagnetic Field Polarity Reversals into the Climate Variability. In Proceedings of the 7th International Conference Problems of Geocosmos, St. Petersburg, Russia, 6–10 October 2008; pp. 146–151. [Google Scholar]
  140. Usoskin, I.G.; Kovaltsov, G.A. Cosmic Rays and Climate of the Earth Possible Connection. Comptes Rendus Geosci. 2008, 340, 441–450. [Google Scholar] [CrossRef]
  141. Kopper, J.; Papamarinopoulos, S. Human Evolution and Geomagnetism. J. Field Archaeol. 1978, 5, 443–452. [Google Scholar] [CrossRef]
  142. Vizán, H.; Azcuy, C.L.; Tófalo, O.R.; Geuna, S.E.; Amenabar, C.R.; Orgeira, M.J.; Renda, E.M.; Braña Carreño, C.; Panarello, H.; Castro, L. The Blake Geomagnetic Field Event Recorded in a Sequence of Marine and Continental Facies Outcropping in the Coast of Buenos Aires Province, Argentina. J. S. Am. Earth Sci. 2019, 92, 506–522. [Google Scholar] [CrossRef]
  143. Gubbins, D. The distinction between geomagnetic excursions and reversals. Geophys. J. Int. 1999, 137, F1–F3. [Google Scholar] [CrossRef]
  144. Glassmeier, K.H.; Richter, O.; Vogt, J.; Möbus, P.; Schwalb, A. The Sun, geomagnetic polarity transitions, and possible biospheric effects: Review and illustrating model. Int. J. Astrobiol. 2009, 8, 147–159. [Google Scholar] [CrossRef]
  145. Knudsen, M.F.; Riisager, P. Is there a link between Earth’s magnetic field and low-latitude precipitation? Geology 2009, 37, 71–74. [Google Scholar] [CrossRef]
  146. Valet, J.P.; Valladas, H. The Laschamp-Mono lake geomagnetic events and the extinction of Neanderthal: A causal link or a coincidence? Quat. Sci. Rev. 2010, 29, 3887–3893. [Google Scholar] [CrossRef]
  147. Nowaczyk, N.R.; Arz, H.W.; Frank, U.; Kind, J.; Plessen, B. Dynamics of the Laschamp geomagnetic excursion from Black Sea sediments. Earth Planet Sci. Lett. 2012, 351, 54–69. [Google Scholar] [CrossRef]
  148. Lascu, I.; Feinberg, J.; Dorale, J.A.; Cheng, H.; Edwards, R.L. Age of the Laschamp excursion determined by U-Th dating of a speleothem geomagnetic record from North America. Geology 2016, 44, 139–142. [Google Scholar] [CrossRef]
  149. Nami, H.G. Detailed Paleomagnetic Records from Ecuador and New Evidence for the Geomagnetic Field Excursion during the Late Pleistocene-Holocene. Geofis. Int. 2015, 54, 127–148. [Google Scholar]
  150. Chaparro, M.A.E.; Böhnel, H.N.; Byrne, R.; Nowaczyk, N.R.; Molina-Garza, R.S.; Parkand, J.; Negendank, F.W. Palaeomagnetic Secular Variation and Rock-Magnetic Studies of Holocene Sediments from a Maar Lake (Hoya de San Nicolas) in Central Mexico. Geophys. J. Int. 2008, 175, 462–476. [Google Scholar] [CrossRef]
  151. Walker, M.; Johnsen, S.; Rasmussen, S.O.; Popp, T.; Steffensen, J.P.; Gibbard, P.; Hoek, W.; Lowe, J.; Andrews, J.; Björck, S.; et al. Formal definition and dating of the GSSP (Global Stratotype Section and Point) for the base of the Holocene using the Greenland NGRIP ice core, and selected auxiliary records. J. Quat. Sci. 2009, 24, 3–17. [Google Scholar] [CrossRef]
  152. Crain, I.K. Possible Direct Causal Relation between Geomagnetic Reversals and Biological Extinctions. GSA Bull. 1971, 82, 2603. [Google Scholar] [CrossRef]
  153. Raup, D.M. Magnetic reversals and mass extinctions. Nature 1985, 314, 341–343. [Google Scholar] [CrossRef]
  154. Loper, D.E.; McCartney, K.; Buzyna, G. A Model of Correlated Episodicity in Magnetic-Field Reversals, Climate, and Mass Extinctions. J. Geol. 1988, 96, 1–15. [Google Scholar] [CrossRef]
  155. Chou, Y.M.; Jiang, X.; Liu, Q.; Hu, H.M.; Wu, C.C.; Liu, J.; Jiang, Z.; Lee, T.Q.; Wang, C.C.; Song, Y.F.; et al. Multidecadally resolved polarity oscillations during a geomagnetic excursion. Proc. Natl. Acad. Sci. USA 2018, 115, 8913–8918. [Google Scholar] [CrossRef] [Green Version]
  156. Pino, M.; Abarzúa, A.M.; Astorga, G.; Martel-Cea, A.; Cossio-Montecinos, N.; Navarro, R.X.; Lira, M.P.; Labarca, R.; Lecompte, M.A.; Adedeji, V.; et al. Sedimentary record from Patagonia, southern Chile supports cosmic-impact triggering of biomass burning, climate change, and megafaunal extinctions at 12.8 ka. Sci. Rep. 2019, 9, 4413. [Google Scholar] [CrossRef]
  157. Nami, H.G. Paleoindio, Cazadores-Recolectores y Tecnología Lítica en el Extremo Sur de Sudamérica Continental. In Arqueología de Cazadores-Recolectores. Límites, Casos y Aperturas. Arq. Cont. 1994, 5, 89–103. [Google Scholar]
  158. Nami, H.G. Theoretical and Epistemological Thoughts on Archaeology and Experimental Lithic Technology. J. Res. Philos. Hist. 2018, 1, 139. [Google Scholar] [CrossRef]
  159. Willingham, D.B.; Nissen, M.J.; Bullemer, P. On the development of procedural knowledge. J. Exp. Psychol. Learn. Mem. Cogn. 1989, 15, 1047–1060. [Google Scholar] [CrossRef]
  160. Squire, L.R. Memory systems of the brain: A brief history and current perspective. Neurobiol. Learn. Mem. 2004, 82, 171–177. [Google Scholar] [CrossRef]
  161. Schiffer, M.B.; Skibo, J.M. Theory and Experiment in the Study of Technological Change. Curr. Anthr. 1987, 28, 595–622. [Google Scholar] [CrossRef]
  162. Nami, H.G. Investigaciones actualísticas para discutir aspectos técnicos de los cazadores-recolectores del tardiglacial: El problema Clovis-Cueva Fell. An. Inst. Patagon. 1997, 25, 152–186. [Google Scholar]
  163. Nami, H.G. Artefactos Bifaciales de los Niveles Inferiores del Alero Cárdenas. In Contribución a la Arqueología del Río Pinturas Provincia de Santa Cruz; Gradin, C.J., Aguerre, A.M., Eds.; Búsqueda de Ayllu: Concepción del Uruguay, Argentina, 1994; pp. 134–151. [Google Scholar]
  164. Nami, H.G. Comentarios y observaciones sobre las puntas de proyectil de los niveles inferiores de La Martita (Provincia de Santa Cruz). In Arqueología y Paleoambiente en la Patagonia Santacruceña Argentina; Aguerre, A.M., Ed.; Búsqueda de Ayllu: Buenos Aires, Argentina, 2003; pp. 65–72. [Google Scholar]
  165. Lahren, L.; Bonnichsen, R. Bone Foreshafts from a Clovis burial in Southwestern Montana. Science 1974, 11, 147–150. [Google Scholar] [CrossRef]
  166. Posth, C.; Nakatsuka, N.; Lazaridis, I.; Skoglund, P.; Mallick, S.; Lamnidis, T.C.; Rohland, N.; Nägele, K.; Adamski, N.; Bertolini, E.; et al. Reconstructing the Deep Population History of Central and South America. Cell 2018, 175, 1185–1197. [Google Scholar] [CrossRef]
  167. Dillehay, T.D. Monte Verde: A Late Pleistocene Settlement in Chile. In Volume II: The Archaeological Context; Smithsonian Institution Press: Washington, DC, USA, 1997. [Google Scholar]
  168. Dillehay, T.D.; Ocampo, C.; Saavedra, J.; Sawakuchi, A.O.; Vega, R.M.; Pino, M.; Collins, M.B.; Cummings, L.S.; Arregui, I.; Villagran, X.S.; et al. Correction: New Archaeological Evidence for an Early Human Presence at Monte Verde, Chile. PLoS ONE 2015, 10, e0145471. [Google Scholar] [CrossRef]
  169. Dillehay, T.D.; Goodbred, S.; Pino, M.; Sánchez, V.F.V.; Tham, T.R.; Adovasio, J.; Collins, M.B.; Netherly, P.J.; Hastorf, C.A.; Chiou, K.L.; et al. Simple technologies and diverse food strategies of the Late Pleistocene and Early Holocene at Huaca Prieta, Coastal Peru. Sci. Adv. 2017, 3, e1602778. [Google Scholar] [CrossRef]
  170. Halligan, J.J.; Waters, M.R.; Perrotti, A.; Owens, I.J.; Feinberg, J.M.; Bourne, M.D.; Fenerty, B.; Winsborough, B.; Carlson, D.; Fisher, D.C.; et al. Pre-Clovis occupation 14,550 years ago at the Page-Ladson site, Florida, and the peopling of the Americas. Sci. Adv. 2016, 2, e1600375. [Google Scholar] [CrossRef]
  171. Waters, M.R.; Keene, J.L.; Forman, S.L.; Prewitt, E.R.; Carlson, D.L.; Wiederhold, J.E. Pre-Clovis projectile points at the Debra L. Friedkin site, Texas—Implications for the Late Pleistocene peopling of the Americas. Sci. Adv. 2018, 4, eaat4505. [Google Scholar] [CrossRef]
  172. Waters, M.R.; Stafford, T.W. Redefining the Age of Clovis: Implications for the Peopling of the Americas. Science 2007, 315, 1122–1126. [Google Scholar] [CrossRef]
  173. Maggard, G.J.; Dillehay, T.D. El Palto Phase (13800–9800 BP). In From Foraging to Farming in the Andes: New Perspectives on Food Production and Social Organization; Dillehay, T., Ed.; Cambridge University Press: Cambridge, UK, 2011; pp. 77–94. [Google Scholar]
  174. Morrow, J.E.; Morrow, T.A. Geographic Variation in Fluted Projectile Points: A Hemispheric Perspective. Am. Antiq. 1999, 64, 215–230. [Google Scholar] [CrossRef]
  175. Politis, G. Fishtail Projectile Points in the Southern Cone of South America: An Overview. In Clovis, Origins and Adaptations; Bonnichsen, R., Turnmire, K.L., Eds.; Center for the Study of the First Americans: Corvallis, OR, USA, 1991; pp. 287–302. [Google Scholar]
  176. Borrero, L.A. Paleoindians without Mammoths and Archaeologists without Projectile Points? The archaeology of the first inhabitants of the Americas. In Paleoindian Archaeology: A Hemispherical Perspective; Morrow, J.E., Gnecco, C., Eds.; University of Florida Press: Gainesville, FL, USA, 2006; pp. 9–20. [Google Scholar]
  177. Nami, H.G. Experiments to Understand North and South American Late Pleistocene Lithic Reduction Sequences: An Actualistic and Comparative Study. In Experiment and Interpretation of Traditional Technologies: Essays in Honor of Errett Callahan; Nami, H.G., Ed.; 2010; pp. 203–225. Available online: https://www.academia.edu/989749/Experiments_to_Replicate_North_and_South_American_Late_Pleistocene_Lithic_Reduction_Sequences_An_Actualistic_and_Comparative_Study (accessed on 12 August 2019).
  178. Miotti, L.L. Patagonia: A paradox for building images of the first Americans during the Pleistocene/Holocene Transition. Quat. Int. 2003, 109, 147–173. [Google Scholar] [CrossRef]
  179. Dillehay, T.D. Profiles in Pleistocene prehistory. In Handbook of South American Archaeology; Silverman, H., Isbell, W.H., Eds.; Springer: New York, NY, USA, 2008; pp. 29–44. [Google Scholar]
  180. Goebel, T.; Waters, M.R.; O’Rourke, D.H. The Late Pleistocene Dispersal of Modern Humans in the Americas. Science 2008, 319, 1497–1502. [Google Scholar] [CrossRef] [Green Version]
  181. O’Brien, M.J. Setting the Stage: The Late Pleistocene Colonization of North America. Quaternary 2019, 2, 1. [Google Scholar] [CrossRef]
  182. Erlandson, J.M.; Braje, T.J. Stemmed Points, the Coastal Migration Theory, and the Peopling of the Americas. In Mobility and Ancient Society in Asia and the Americas; Frachetti, M., Spengler, R., Eds.; Springer: Cham, Switzerland, 2015; pp. 49–58. [Google Scholar]
  183. Pearson, G.A. Pan-American Paleoindian Dispersals and the Origins of Fishtail Projectile Points as Seen through the Lithic Raw-Material Reduction Strategies and Tool-Manufacturing Techniques at the Guardiría Site, Turrialba Valley, Costa Rica. In The Settlement of the American Continents; Barton, C.M., Clark, G.A., Yesner, D.R., Pearson, G.A., Eds.; The University of Arizona Press: Tucson, AZ, USA, 2004; pp. 85–102. [Google Scholar]
  184. Faught, M.K. Paleoindian Archaeology in Florida and Panama. In Paleoindian Archaeology: A Hemispherical Perspective; Morrow, J.E., Gnecco, C., Eds.; University of Florida Press: Gainesville, FL, USA, 2006; pp. 164–218. [Google Scholar]
  185. Nami, H.G. Paleo American finds from Venezuela: Evidence to discuss the spread of Fell Points and the peopling of Northern South America. Rev. Cad. Ceom 2016, 29, 212. [Google Scholar] [CrossRef]
  186. Nami, H.G. Fishtailed Projectile Points in the New World: Remarks and Hypothesis on the Peopling of Northern South America and Beyond. 2019; in preparation. [Google Scholar]
  187. Miotti, L. La fachada atlántica, como puerta de ingreso alternativa de la colonización humana de América del Sur durante la transición Pleistoceno/Holoceno. In 2 Simposio Internacional del Hombre Temprano en América; Jiménez López, J.C., Ed.; Instituto Nacional de Antropología e Historia: Mexico City, Mexico, 2006; pp. 155–188. [Google Scholar]
  188. Grosjean, M.; Nuñez, L.; Cartagena, I. Palaeoindian Occupation of the Atacama Desert, northern Chile. J. Quat. Sci. 2005, 20, 643–653. [Google Scholar] [CrossRef]
  189. Jackson, D.; Méndez, C.; Seguel, R.; Maldonado, A.; Vargas, G. Initial Occupation of the Pacific Coast of Chile during Late Pleistocene Times. Curr. Anthr. 2007, 48, 725–731. [Google Scholar] [CrossRef] [Green Version]
  190. Méndez, C.; Delaunay, A.N.; Seguel, R.; Maldonado, A.; Murillo, I.; Jackson, D.; Aspillaga, E.; Izaurieta, R.; Méndez, V.; Fernández, M. Late Pleistocene to early Holocene high-quality quartz crystal procurement from the Valiente quarry workshop site (32°S, Chile, South America). PLoS ONE 2018, 13, e0208062. [Google Scholar] [CrossRef]
  191. Nuñez, L.; Casamiquela, R.; Schiappacasse, V.; Niemeyer, H.; Villagrán, C. Cuenca de Taguatagua en Chile: El ambiente del Pleistoceno y ocupaciones humanas. Rev. Chil. Hist. Nat. 1994, 67, 503–519. [Google Scholar]
  192. Nami, H.G. Nuevos datos en relación a las puntas de proyectil Paleoindias encontradas en el Cono Sur (Neuquén, Argentina). Palimpsesto Rev. Arqueol. 1992, 1, 71–74. [Google Scholar]
  193. Aráoz, C.J.P.; Nami, H.G. The First Paleoindian Fishtail Point Find in Salta Province, Northwestern Argentina. Archaeol. Discov. 2014, 2, 26–30. [Google Scholar] [CrossRef] [Green Version]
  194. Nami, H.G. Fell Point Technology from Mendoza Province in Western Argentina. Paleo Am. 2018, 4, 251–257. [Google Scholar] [CrossRef]
  195. Weitzel, C.; Mazzia, N.; Flegenheimer, N. Assessing Fishtail points distribution in the southern Cone. Quat. Int. 2018, 473, 161–172. [Google Scholar] [CrossRef]
Figure 1. Locations maps of the studied area. (a) South America, (b) southern Patagonia indicating the Cerro Benitez area with an arrow, (c) its location close to the Ultima Esperanza Sound. References: CB: Benitez Hill, LS: Lago Sofía, UES: Ultima Esperanza Sound. Modified after Google and NASA maps.
Figure 1. Locations maps of the studied area. (a) South America, (b) southern Patagonia indicating the Cerro Benitez area with an arrow, (c) its location close to the Ultima Esperanza Sound. References: CB: Benitez Hill, LS: Lago Sofía, UES: Ultima Esperanza Sound. Modified after Google and NASA maps.
Quaternary 02 00028 g001
Figure 2. Location of the Benitez Hill and Ultima Esperanza Sound in a schematic geomorphological profile from the Monument Moore hill (west) to Sierra Dorotea on the Chile–Argentina border (east). Graphic design in Figures 2, 3, 6, 7 and 18: H. G. Nami and G. Paez Reina.
Figure 2. Location of the Benitez Hill and Ultima Esperanza Sound in a schematic geomorphological profile from the Monument Moore hill (west) to Sierra Dorotea on the Chile–Argentina border (east). Graphic design in Figures 2, 3, 6, 7 and 18: H. G. Nami and G. Paez Reina.
Quaternary 02 00028 g002
Figure 3. Location maps of the main Late Pleistocene paleontological and archaeological cave sites in the area. (a) The Señoret and Benitez hills. 1. Lago Sofía 1 and 2, 2. de la Ventana, 3. Mylodon, 4. del Medio, 5. Chica. The arrow on the map of South America shows the Ultima Esperanza region, (b) detail of the SW slope of the Benitez Hill with the locations of the Mylodon (CMi), Medio (CM), and Chica (CC) caves in relation to the “Silla del Diablo” (SD) rock formation in the Mylodon Cave National Park.
Figure 3. Location maps of the main Late Pleistocene paleontological and archaeological cave sites in the area. (a) The Señoret and Benitez hills. 1. Lago Sofía 1 and 2, 2. de la Ventana, 3. Mylodon, 4. del Medio, 5. Chica. The arrow on the map of South America shows the Ultima Esperanza region, (b) detail of the SW slope of the Benitez Hill with the locations of the Mylodon (CMi), Medio (CM), and Chica (CC) caves in relation to the “Silla del Diablo” (SD) rock formation in the Mylodon Cave National Park.
Quaternary 02 00028 g003
Figure 4. General views of the Cerro Benitez (a), and Cueva del Medio (bd). (b) entrance, (c,d) frontal and inner portions. Except when clearly stated, all the photographs and drawings are by Hugo G. Nami.
Figure 4. General views of the Cerro Benitez (a), and Cueva del Medio (bd). (b) entrance, (c,d) frontal and inner portions. Except when clearly stated, all the photographs and drawings are by Hugo G. Nami.
Quaternary 02 00028 g004
Figure 5. Images of the excavation made at CM during the different field-works. (a) Made during the first expedition. Several looters’ pits can be seen around it, one of which was used to start the excavation, (b) Paleo-American level exposed during the third expedition.
Figure 5. Images of the excavation made at CM during the different field-works. (a) Made during the first expedition. Several looters’ pits can be seen around it, one of which was used to start the excavation, (b) Paleo-American level exposed during the third expedition.
Quaternary 02 00028 g005
Figure 6. Map of the CM showing the locations of the author’s excavations.
Figure 6. Map of the CM showing the locations of the author’s excavations.
Quaternary 02 00028 g006
Figure 7. Examples of stratigraphic sections from the excavations of CM. (ac) East/west direction, (d) north/south direction. References: 0 = Datum, the numbers in the upper part of the datum indicate the grids; the rectangles at the bottom show the layers. The black lenses in (b,c) are carbonized sand and a hearth, respectively. The arrows indicate the position of the PA archaeological level at the top of layer four.
Figure 7. Examples of stratigraphic sections from the excavations of CM. (ac) East/west direction, (d) north/south direction. References: 0 = Datum, the numbers in the upper part of the datum indicate the grids; the rectangles at the bottom show the layers. The black lenses in (b,c) are carbonized sand and a hearth, respectively. The arrows indicate the position of the PA archaeological level at the top of layer four.
Quaternary 02 00028 g007
Figure 8. Frontal (a,b) and upper (c,d) views of one of the hearths found at CM, Hippidion saldiasi bones and a Fell point indicated by the arrow and red circles in (b,d). Note the pebble layer sealing the archaeological remains pointed with yellow arrows.
Figure 8. Frontal (a,b) and upper (c,d) views of one of the hearths found at CM, Hippidion saldiasi bones and a Fell point indicated by the arrow and red circles in (b,d). Note the pebble layer sealing the archaeological remains pointed with yellow arrows.
Quaternary 02 00028 g008
Figure 9. Images of a pile of bones placed (a) over a typical PA hearth with a concave base and a mandible of an infant horse placed on it (b).
Figure 9. Images of a pile of bones placed (a) over a typical PA hearth with a concave base and a mandible of an infant horse placed on it (b).
Quaternary 02 00028 g009
Figure 10. Relationship between the percentages of completeness (% Comp.) and the minimum number of individuals (MNI)/minimum number of skeletal elements (MNE) according to the different taxa exhumed in layer four. The rectangles show the consumed species (modified after Figure 6.2.7 of [31]). References—1: Lama Guanicoe and Lama cf. owenii, 2: Hippidion saldiasi, 3: Mylodon cf. dawinii, 4: Canidae (enclose at least Dusicyum avus).
Figure 10. Relationship between the percentages of completeness (% Comp.) and the minimum number of individuals (MNI)/minimum number of skeletal elements (MNE) according to the different taxa exhumed in layer four. The rectangles show the consumed species (modified after Figure 6.2.7 of [31]). References—1: Lama Guanicoe and Lama cf. owenii, 2: Hippidion saldiasi, 3: Mylodon cf. dawinii, 4: Canidae (enclose at least Dusicyum avus).
Quaternary 02 00028 g010
Figure 11. Plot of the calibrated dates obtained from the Pleistocene deposit from CM. The highlighted frame shows the time-span for early human occupation. The green arrows points to the dates obtained from samples coming from the PA occupation.
Figure 11. Plot of the calibrated dates obtained from the Pleistocene deposit from CM. The highlighted frame shows the time-span for early human occupation. The green arrows points to the dates obtained from samples coming from the PA occupation.
Quaternary 02 00028 g011
Figure 12. Distribution of the PA artifacts in the main excavation. The detail of each piece is given in Table 2.
Figure 12. Distribution of the PA artifacts in the main excavation. The detail of each piece is given in Table 2.
Quaternary 02 00028 g012
Figure 13. Fell points found in the Paleoamerican level from CM. (a,b) from Nami’s excavation, (c) approximate drawing based on photograph after Figure 14 of [29], (d) from the Epullán locality.
Figure 13. Fell points found in the Paleoamerican level from CM. (a,b) from Nami’s excavation, (c) approximate drawing based on photograph after Figure 14 of [29], (d) from the Epullán locality.
Quaternary 02 00028 g013
Figure 14. Bifacial and unifacial implements exhumed in the Paleo-American level of CM (a,b) “Fishtail” projectile points, (c,d) unifacial tools.
Figure 14. Bifacial and unifacial implements exhumed in the Paleo-American level of CM (a,b) “Fishtail” projectile points, (c,d) unifacial tools.
Quaternary 02 00028 g014
Figure 15. Unifacial tools coming from the Paleo-american level of CM. Depictions (a–d) are given in text and Table 2.
Figure 15. Unifacial tools coming from the Paleo-american level of CM. Depictions (a–d) are given in text and Table 2.
Quaternary 02 00028 g015
Figure 16. Drawings illustrating diverse Paleo-American flaked artifacts. (ae) from the CM, (fg) Fell’s cave. (a,c–g) Unifacial tools, (b) early manufacturing stage of a “fisthail” point.
Figure 16. Drawings illustrating diverse Paleo-American flaked artifacts. (ae) from the CM, (fg) Fell’s cave. (a,c–g) Unifacial tools, (b) early manufacturing stage of a “fisthail” point.
Quaternary 02 00028 g016
Figure 17. Map of the ecological regions in South America at the Pleistocene–Holocene transition (based on [106,107], and modified after Figure 2 of [109].
Figure 17. Map of the ecological regions in South America at the Pleistocene–Holocene transition (based on [106,107], and modified after Figure 2 of [109].
Quaternary 02 00028 g017
Figure 18. World map and location of the virtual geomagnetic poles calculated from samples with anomalous directions recorded during the Pleistocene-Holocene transition and early Holocene in some sites from: (a) Chile, (b) Argentina, (c) Ecuador, (d) South Africa, (f) The totality of Virtual Geomagnetic Poles from the locales illustrated in (ad). Modified after [129,136,137,149].
Figure 18. World map and location of the virtual geomagnetic poles calculated from samples with anomalous directions recorded during the Pleistocene-Holocene transition and early Holocene in some sites from: (a) Chile, (b) Argentina, (c) Ecuador, (d) South Africa, (f) The totality of Virtual Geomagnetic Poles from the locales illustrated in (ad). Modified after [129,136,137,149].
Quaternary 02 00028 g018
Table 1. List of AMS dates corresponding to the Pleistocene–Holocene transition obtained at CM. The dates in bold fonts comes from a level with mix of triangular points.
Table 1. List of AMS dates corresponding to the Pleistocene–Holocene transition obtained at CM. The dates in bold fonts comes from a level with mix of triangular points.
Material DatedOriginLab. Id.DateCalibrated RangeReference
Charcoal29/8d–29/7cPITT–03449595 ± 11510,595–11,215[8]
Mammal bones18/11Beta–402819770 ± 7010,809–11,329
Charcoal23/5abGr–N1491210,310 ± 7011,825–12,404
Burn Mammal bones27/10d–26/9aBeta–5810510,350 ± 13011,647–12,592[24]
Charcoal38/40Beta 31953810,410 ± 8012,073–12,524[29]
7/19–8/19Beta-5252210,430 ± 8012,039–12,564[16]
Camelidae bone27/10d–26/9aNUTA-173410,430 ± 10011,988–12,619[24]
L. guanicoe bone28/6cNUTA-173510,450 ± 10012,020–12,637
Burn Mammal bones27/10d–26/9aGr-N 1491110,550 ± 12012,099–12,709[8]
L. guanicoe boneIndet.Oxa–2196010,630 ± 5512,435–12,710[61]
Hippidion saldiasi bone53/34Beta 34442810,680 ± 4012,571–12,711[29]
29/8cNUTA-181110,710 ± 10012,421–12,767[24]
L. guanicoe bone28/6bNUTA-233210,710 ± 19012,077–13,017
Canidae bone54/35Beta 34190310,710 ± 5012,577–12,724[29]
L. guanicoe boneIndet.Oxa–2147010,720 ± 4512,587–12,728[61]
Indet.Oxa-2138510,725 ± 4512,591–12,730
Oxa-2147210,740 ± 5012,594–12,737
Oxa-2147110,750 ± 5012,599–12,737
Hippidion saldiasi bone Oxa-2612510,810 ± 4512,668–12,766
Camelidae bone28/6dNUTA-181210,850 ± 13012,560–13,037[24]
Felidae, cf. Panthera54/35Beta 34443010,860 ± 4012,692–12,798[29]
Hippidion saldiasi bone28/5aAA–10023510,860 ± 11012,595–13,008
Hippidion saldiasi bone27/10d–26/9aNUTA–233110.860 ± 16012.432–13.084[24]
Mammal boneNon dataAA-1301810,885 ± 9012,672–12,996[35]
L. gracilis boneIndet.Oxa-2146910,925 ± 4512,702–12,897[61]
CharcoalInnerBeta-3908110,930 ± 23012,280–13,300[22]
Camelidae. bone26/5aNUTA-233010,960 ± 15012,639–13,114[24]
27/10d-26/9aNUTA-219711,040 ± 25012,430–13,432
25/7bNUTA-173711,120 ± 13012,726–13,206
Table 2. The salient morphological and technological attributes of the CM lithic artifacts reported in this paper. References: AMR: Artifact with marginal retouch, ARM: Artifact with marginal retouch in a restricted angle, B: Biface, BA: Bifacially flaked artifact, C: Chert, Ch: Chalcedony, EA: Edge angle, FB: Fractured biface, FUA: Fragment of unifacial artifact, FSR: Flake with summary retouch, L: Length, PP: Projectile point, PS: Perimeter scraper, R: Raclette, S: Scraper, SS: Side scraper, Tu: Tuff, T: Thickness, W: Width. Note: All the measurements are given in mm, (): indicates a fractured piece.
Table 2. The salient morphological and technological attributes of the CM lithic artifacts reported in this paper. References: AMR: Artifact with marginal retouch, ARM: Artifact with marginal retouch in a restricted angle, B: Biface, BA: Bifacially flaked artifact, C: Chert, Ch: Chalcedony, EA: Edge angle, FB: Fractured biface, FUA: Fragment of unifacial artifact, FSR: Flake with summary retouch, L: Length, PP: Projectile point, PS: Perimeter scraper, R: Raclette, S: Scraper, SS: Side scraper, Tu: Tuff, T: Thickness, W: Width. Note: All the measurements are given in mm, (): indicates a fractured piece.
Piece #GridDepthToolMaterialBlankLWTFigure
-18/10a−27SSTuF785014-
-25/5a–c−35/−40PP B4018813a, 14a
14128/5c−18SS F37437-
17228/5a−21SS 925016-
-26/6a, d−45BA 3532816b
7023/6b−45PSCh 433513-
18729/6c−46FSR 898636-
33130/6a−40K (59)38815d
37830/6d−40BACh 50398-
225/7b−25/−30ARMCInd(28)351216a
6925/7d−30/−35BA (31)(15)8-
18928/7b−53STuF69501514c
16028/7c−57SS 38534-
7625/8a−69SS TN904619-
19628/8c−74K F543216-
7325/9a−71ARM 100657-
7025/9b−68SS (60)8410-
-25/9c−71SS (50)(50)6
7525/9b−70/−75SS (28)(40)38-
6925/9c−71FUA F(47)(47)7-
-26/9a−86PP B4827613b, 14b
19827/9b−83K 7532814d, 16d
-27/9c−92R F8231316e
20428/9c−83AMR (53)3512-
38128/9c−80/−85PS 62531115b
-27/11b110/−115SS 738221-
21028/11a−132SS Ind.93752316c
34829/11b−108SS 102321215c
38330/11d−113SS 859525-
36529/12d−115FB 364211-
36129/12b−123SS F56(37)10-

Share and Cite

MDPI and ACS Style

Nami, H.G. Paleoamerican Occupation, Stone Tools from the Cueva del Medio, and Considerations for the Late Pleistocene Archaeology in Southern South America. Quaternary 2019, 2, 28. https://0-doi-org.brum.beds.ac.uk/10.3390/quat2030028

AMA Style

Nami HG. Paleoamerican Occupation, Stone Tools from the Cueva del Medio, and Considerations for the Late Pleistocene Archaeology in Southern South America. Quaternary. 2019; 2(3):28. https://0-doi-org.brum.beds.ac.uk/10.3390/quat2030028

Chicago/Turabian Style

Nami, Hugo G. 2019. "Paleoamerican Occupation, Stone Tools from the Cueva del Medio, and Considerations for the Late Pleistocene Archaeology in Southern South America" Quaternary 2, no. 3: 28. https://0-doi-org.brum.beds.ac.uk/10.3390/quat2030028

Article Metrics

Back to TopTop