Manuscript Click here to access/download;Manuscript;Text.docx

Click here to view linked References The technocomplex of the Iberian Atlantic margin as an example of technology continuity through the Middle

1 2 3 4 5 Eduardo Méndez-Quintas 1,2*; Manuel Santonja 3,2; Lee J. Arnold 4; João Pedro Cunha-Ribeiro 5; Pedro Xavier 6 6 4 7 7 6 7 da Silva ; Martina Demuro ; Mathieu Duval ; Alberto Gomes ; José Meireles ; Sérgio Monteiro-Rodrigues 9 2 8 ; Alfredo Pérez-González 9 10 11 1 Grupo de Estudos de Arqueoloxía, Antigüidade e Territorio (GEAAT), University of Vigo, Campus As Lagoas, 32004 12 13 Ourense, Spain. 14 2 IDEA (Instituto de Evolución en África), University of Alcalá de Henares, Covarrubias 36, 28010 Madrid, Spain 15 16 3 Centro Nacional de Investigación sobre la Evolución Humana (CENIEH). Paseo de Atapuerca, 3. 09002 Burgos, 17 España. 18 19 4 School of Physical Sciences, Environment Institute, and Institute for Photonics and Advanced Sensing (IPAS), 20 University of Adelaide, North Terrace Campus, Adelaide SA 5005, Australia. 21 22 5 Faculdade de Letras da Universidade de Lisboa, Centro de Arqueologia da Universidade de Lisboa (UNIARCH) e 23 Lab2pt, 1600-214 Lisbon, Portugal. 24 25 6 Universidade do Minho; Lab2pt, 39 av. Central, 4700-Braga, Portugal 26 7 27 Australian Research Centre for Human Evolution. Environmental Futures Research Institute, Griffith University, 170 28 Kessels Road Nathan, QLD 4111, Australia.

29 8 30 Faculdade de Letras, Universidade do Porto; Centro de Estudos de Geografia e Ordenamento do Território (CEGOT). 31 Via Panorâmica s/n, P 4150-564, Porto, Portugal 32 9 Faculdade de Letras, Universidade do Porto; Centro de Investigação Transdisciplinar “Cultura, Espaço e Memória” 33 34 (CITCEM). Via Panorâmica s/n, P 4150-564, Porto, Portugal 35 36 37 * Corresponding author. E-mail: [email protected] 38 39 40 41 Eduardo Méndez-Quintas: 0000-0001-8272-873X 42 43 Manuel Santonja: 0000-0001-5848-9011 44 45 Lee J. Arnold: 0000-0001-9603-3824 46 47 João Pedro Cunha-Ribeiro: 0000-0003-2614-3555 48 49 Pedro Xavier da Silva: 0000-0002-4217-0126 50 51 Martina Demuro: 0000-0002-5799-4216 52 Mathieu Duval: 0000-0003-3181-7753 53 54 Alberto Gomes: 0000-0001-8600-0383 55 56 José Meireles: 0000-0002-8416-0591 57 58 Sérgio Monteiro-Rodrigues: 0000-0002-2532-3602 59 60 Alfredo Pérez-González: 0000-0003-1122-9313 61 62 1 63 64 65 Abstract This article provides a synthesis of the Middle Pleistocene hominin record of West Iberia, which comprises sites displaying abundant concentrations of large flake Acheulean (LFA) assemblages, as well as isolated 1 2 examples of Early Middle Palaeolithic (EMP) technology. These sites typically have age ranges spanning 3 4 Marine Isotopic Stages (MIS) 11-6, within the second half of the Middle Pleistocene, and are primarily located 5 in fluvial environments related to the main regional river basins. The LFA sites display extensive occurrences 6 7 of handaxes and cleavers on flake blanks (detached from large cores), with a large number of knapping 8 9 remains, such as flakes or small-medium cores, showing elementary reduction patterns. Over the identified 10 age range of these sites, especially during the MIS 9-6 interval, we observe constant technological stability, 11 12 without strong variations over time, and independent of the functionality of individual sites. These fixed 13 14 technological and behavioural patterns reinforce the African affinities of the southwestern European 15 Acheulean, in contrast to Acheulean assemblages identified in the northernmost areas of Europe. 16 17 18 Keywords 19 20 Middle Pleistocene; Lower Palaeolithic; Large Flake Acheulean (LFA); Early Middle Palaeolithic (EMP); West 21 22 Iberia 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 2 63 64 65 INTRODUCTION

The Iberian Peninsula is an important region for understanding the biological and cultural processes that 1 shaped human history in Europe during the Lower and Middle Pleistocene (~1.4 to 0.125 million years ago 2 3 (Ma)). The presence of numerous archaeological sites from the Guadix-Baza basin (Toro-Moyano et al. 2011; 4 Toro-Moyano et al. 2013) and the Atapuerca complex (Carbonell et al. 1995; Carbonell et al. 2008; Ollé et 5 6 al. 2013), as well as those of La Boella (Vallverdu et al. 2014), Ambrona (Falguères et al. 2006; Santonja et 7 8 al. 2017), Torralba (Sánchez-Cervera et al. 2015), Áridos (Santonja et al. 1980), Cuesta de la Bajada 9 (Santonja et al. 2014) and Cueva de Bolomor (Fernández Peris et al. 2012), have provided important 10 11 information in this context. Most of these sites are located in the interior of the Iberian Peninsula, and, until 12 13 recently, relatively little was known about the age or representative industry assemblages of in situ sites 14 located on the western side of the Iberian Peninsula. Over the last few decades, the expansion of 15 16 archaeological research into different parts of the Iberian Atlantic margin (such as the Miño River and Lis 17 18 River basins), and the detailed examination of previously known sites across the region (such as in the lower 19 Tagus basin), has significantly increased the scope and coverage of the Iberian Middle Pleistocene 20 21 archaeological record (Daura et al. 2017; Méndez-Quintas et al. 2019; Méndez-Quintas et al. 2018b; Proença 22 23 Cunha et al. 2017a; Proença Cunha et al. 2017b; Texier et al. 1995). A synthesis of these records and 24 examination of their regional significance is therefore timely. 25 26 27 The cultural and biological changes associated with the Middle Pleistocene human occupation of Europe is 28 29 a focus of constant debate (Antoine et al. 2019; Arsuaga et al. 2014; Bermúdez de Castro and Martinón- 30 31 Torres 2013; Bermúdez de Castro et al. 2016; Dennell et al. 2011; Harvati et al. 2019; Hérisson et al. 2016a; 32 Muttoni et al. 2018; Rocca 2016; Santonja et al. 2016; Soriano and Villa 2017; Villa et al. 2016). The role of 33 34 the Acheulean technocomplex is especially important for these ongoing debates, as it shows strong variability 35 36 throughout Europe, both in terms of its chronological development and its basic technological features 37 (Antoine et al. 2019; Pereira et al. 2018; Santonja et al. 2016; Santonja and Villa 2006). In this paper we 38 39 provide an overview of the contributions made to these discussions by archaeological research undertaken 40 41 on the western side of the Iberian Peninsula, which includes extensive examples of sites with large flake 42 Acheulean assemblages (LFA sensu Sharon 2010) spanning the second half of the Middle Pleistocene. 43 44 45 CHRONOLOGICAL FRAMEWORK 46 47 48 The first human occupation of the Iberian Peninsula occurred more than 1 Ma (Carbonell et al. 2008; Martínez 49 et al. 2010; Toro-Moyano et al. 2013) and was characterized by non-Acheulean core and flake industries 50 51 (Mosquera et al. 2018; Toro-Moyano et al. 2011). Some researchers consider that the first Acheulean 52 53 presence in the region is registered at the site of La Boella (Tarragona) (but see discussions in Santonja et 54 al., 2016 and Méndez Quintas et al., 2018), which has been dated to between 1.0 and 0.8 Ma (Vallverdu et 55 56 al. 2014). Further examples of sites that may have analogous chronologies (although without numerical 57 58 dating) include several small lithic assemblages associated with fluvial terraces at heights of more than +40 59 m in the large river basins of the Iberian plateau (Rubio-Jara et al. 2016; Santonja and Pérez-González 2010); 60 61 62 3 63 64 65 some of which contain elements that would suggest an Acheulean attribution. The majority of Lower Palaeolithic sites in the Iberian Peninsula (those containing Acheulean and Early Middle Palaeolithic (EMP) implements) are dated to the Middle Pleistocene, emerging mainly after Marine Isotopic Stage (MIS) 13 and 1 2 prevailing until MIS 6, or even MIS 5 (Méndez-Quintas et al. 2019; Proença Cunha et al. 2017b; Rubio-Jara 3 et al. 2016; Sánchez-Cervera et al. 2015). In this interval of around 400 ka, the number of sites increases 4 5 exponentially over time (Santonja and Villa 2006; Turq et al. 2010). The majority of these sites are associated 6 7 with the middle terraces of Atlantic basins in the Iberian hinterland (Miño, Duero, Tajo, Lis, Guadiana and 8 Guadalquivir), which exhibit elevations of +40-20 m (Cunha Ribeiro 1999; Rubio-Jara et al. 2016; Santonja 9 10 and Pérez-González 2010; Santonja and Villa 2006), as well as a smaller number of cave and open-air sites 11 12 not directly associated with specific fluvial terrace heights (Santonja and Pérez-González 2010; Santonja and 13 Villa 2006). Dating of the archaeological deposits in these hinterland Iberian basins confirms an age range 14 15 of 450-180 thousand years ago (ka) for the expansion of the Acheulean tradition in the region (Martins et al. 16 17 2010; Moreno et al. 2012; Rosina et al. 2014; Silva et al. 2017). Within this same age range, other European 18 regions contain sites displaying EMP industries, which are technologically different and not directly related to 19 20 the LFA (Hérisson et al. 2016b; Locht et al. 2016; Santonja et al. 2016; Turq et al. 2010). 21 22 23 Over recent years, an extensive volume of chronological data (mainly luminescence and electron spin 24 resonance (ESR) ages) has been produced for the middle fluvial terraces of the lower Atlantic rivers (Cunha 25 26 et al. 2008; Daura et al. 2017; Demuro, Arnold, Duval, Méndez-Quintas, et al. 2020; Demuro, Arnold, Duval, 27 28 Spooner, et al. 2020; Méndez-Quintas et al. 2019; Méndez-Quintas et al. 2018b; Proença Cunha et al. 2017b; 29 30 Ramos et al. 2012; Rosina et al. 2014; Silva et al. 2017), providing a strong chronological framework and 31 placing the associated Acheulean sites in the second half of the Middle Pleistocene, between MIS 11 and 32 33 MIS 6 (~424-130 ka; MIS boundaries are from Lisiecki and Raymo, 2005), and mainly between MIS 9 and 34 35 MIS 6 (~337-130 ka) (Fig. 1). 36 37 Archaic pre-Acheulean lithic industry has also been previously reported from high terraces along the coast 38 39 of the Portuguese Extremadura (Açafora, Magoito, Seixosa, Praia da Aguda, Alto do Leião) and the Miño 40 41 River basin, which represent marine or fluvial sedimentary contexts with inferred Lower Pleistocene or earlier 42 ages (Cano Pan et al. 1999; Cano Pan et al. 1997; Raposo 1985; Raposo and Carrera 1986). These sites 43 44 have limited assemblages, uncertain lithic industry attributions, and they are generally from problematic and 45 46 undated sedimentary context. As such, we take a prudent position with the interpretation of these sites, and, 47 based on current information, we do not consider there to be sufficiently secure evidence for human presence 48 49 prior to the Middle Pleistocene in the Iberian Atlantic margin (Raposo and Cardoso 2000; Raposo and 50 51 Santonja 1995). 52 53 MIDDLE PLEISTOCENE ARCHAEO-PALAEONTOLOGICAL RECORD IN WESTERN IBERIA 54 55 56 The known Middle Pleistocene archaeological record from the region has proved to be more important than 57 58 previous syntheses originally suggested (e.g., Chauhan et al. 2017; Moncel et al. 2018), mainly due to the 59 presence of numerous open-air sites, associated with fluvial environments in the larger river basins (Miño, 60 61 62 4 63 64 65 Lis or Tagus rivers), but also due to the presence of important sites from coastal and karstic environments (Fig. 2). However, the predominance of granitic rocks in this region often impedes the conservation of faunal remains and limits, to some extent, the nature of the archaeo-palaeontological records; although there are 1 2 some exceptions (Antunes 1988; Cardoso 1993; Marks et al. 2002a; Sousa and Figueiredo 2001; Vaquero 3 et al. 2017). 4 5 6 7 8 9 The coastline 10 11 12 The existence of Acheulean sites along the Atlantic coastline has been known for many years (Rodríguez 13 Asensio 1983; Zbyszewski 1943). In some cases, assemblages are without stratigraphic context or are found 14 15 in allochthonous positions (e.g., the Louselas site, Spain), which limits reliable age attribution (Álvarez-Alonso 16 17 2014; Ramil Rego and Ramil Soneira 2014). In other cases, excavated sites are associated with marine 18 sediments and continental colluvium deposits of possible Middle Pleistocene age, but these sedimentary 19 20 sequences remain undated. This situation applies to the large number of LFA sites located in the Minho coast 21 22 region (north-western Portugal) (Meireles 1992; Texier et al. 1995). Among these, the Marinho site is the 23 more representative, and includes a large excavated surface (~96 m2) from which 2259 pieces (da Silva 24 25 2012; Meireles 1992) have been recovered. The stratigraphic sequence of the site has been interpreted as 26 27 a marine terrace located +14 m above the current level of the sea, and overlain by several colluvium cycles; 28 the main archaeological level is associated with the lower colluvium layer (da Silva 2012; Meireles 1992). 29 30 The marine terrace is considered to have formed during MIS 7, with the colluvium cycles having an age older 31 32 than MIS 6 (Meireles 1992; Texier et al. 1995). The lithic industry is made from quartzite, and to a lesser 33 extent quartz, sourced from local pebbles and cobbles. The lithic assemblages have abundant knapping 34 35 products (flakes and waste) and cores, with a representative number of flake tools and large cuttings tools 36 37 (LCTs), including handaxes, flake cleavers and trihedral picks (Fig. 3 a-c). The core reduction patterns are 38 simple with monopolar, peripheral or discoidal organization, but without standardized types such as Levallois 39 40 or Quina (da Silva 2012; Meireles 1992). 41 42 43 Around the city of Oporto, several lithic assemblages (some with clear LFA features) have recently been 44 reported in association with Middle Pleistocene marine or continental formations (Monteiro-Rodrigues and 45 46 Cunha Ribeiro 2014; Monteiro-Rodrigues et al. 2016; Monteiro-Rodrigues and González 2010). The Cerro 47 2 48 site (Portugal) is an LFA site that was excavated from 1989 to 1992 and covers an area of 22 m . The 49 stratigraphic sequence of the site is composed of a basal marine pebble layer capped by several colluvium 50 51 levels. The main lithic assemblage is found in the lower colluvium, although some artefacts were also 52 53 recovered from marine sediments (Monteiro-Rodrigues and Cunha Ribeiro 2014). In this collection, the LCTs 54 (mainly handaxes, a few cleavers and other macro-tools) have a high numerical significance, with 127 LCTs 55 56 recovered from a total of 616 artefacts (Fig. 3 g). The main raw material is the local quartzite pebbles and 57 58 cobbles, although quartz blanks are also present. The handaxes usually have pointed morphology and the 59 most frequent blanks are pebbles or large flakes (Monteiro-Rodrigues and Cunha Ribeiro 2014). Another 60 61 62 5 63 64 65 relevant site with Acheulean artefacts is Praia da Aguda (Portugal) (Monteiro-Rodrigues and González 2010). This site has an uncertain stratigraphic position, but it has produced over 120 artefacts, including typical LCTs, shaped on local quartz (Fig. 3 d-f) (Monteiro-Rodrigues and González 2010). This circumstance is 1 2 unique, because quartz is not a common raw material for knapping handaxes or flake cleavers in the Iberian 3 Acheulean, especially when there are other suitable raw materials (i.e., quartzite) in close proximity (Santonja 4 5 and Villa 2006). 6 7 8 Miño River basin 9 10 The Miño basin is the main drainage system for the northern sector of the study area and is an important 11 12 Plio-Pleistocene infill. In the lower basin area, at least 9 fluvial terraces with heights ranging from +108 m to 13 14 +4 m above the current river level have been identified (Méndez Quintas et al. 2017). Available 15 geochronological data from the middle river terrace levels (+40-20 m) indicates a Middle Pleistocene age for 16 17 these fluvial formations, ranging between MIS 9 and MIS 6 (Demuro, Arnold, Duval, Méndez-Quintas, et al. 18 19 2020; Demuro, Arnold, Duval, Spooner, et al. 2020; Méndez-Quintas et al. 2018b; Viveen et al. 2012). 20 21 th 22 The existence of Lower Palaeolithic materials in the area has been known since the beginning of the 20 23 century (Álvarez Blázquez and Bouza Brey 1949; Bouza Brey and Álvarez Blázquez 1954; Viana 1930). 24 25 However, research in the basin did not progress significantly after the discovery of the infamous Gándaras 26 27 de Budiño site (Aguirre 1964; Aguirre and Butzer 1967). Upstream of this site, in the outskirts of the city of 28 Ourense, there are some historical references to decontextualized Acheulean materials (mainly handaxes), 29 30 but few sites are in stratigraphic position or in a sufficient state of conservation to enable reliable interpretation 31 32 (Ameijenda Iglesias et al. 2010; Rodríguez Gracia 1976; Vázquez Varela 1973). 33 34 The main Acheulean sites along the Miño River are located in the lower part of basin and are associated with 35 36 the fluvial staircase, at elevations ranging between +40 m and +20 m above current river level. Sites with 37 38 extensive LFA assemblages include Arbo (Méndez-Quintas et al. 2019; Méndez Quintas et al. 2018), 39 Gándaras de Budiño (Méndez-Quintas 2007, 2008; Méndez-Quintas et al. 2018a), Fillaboa and Porto Maior 40 41 (Méndez-Quintas et al. 2018b) in Spain, and the site of Carvalhas (Cunha Ribeiro et al. 2017) on the 42 43 Portuguese side of the basin. 44 45 The Arbo site is located in a hanging valley and contains a dense concentration of lithic remains made from 46 47 allochthones raw materials. The sedimentary sequence preserved at the Arbo site, which is situated on top 48 49 of the altered feldspar alkaline granite, is composed of a lower thin level of matrix-supported gravels and 50 cobbles (level OC1) that have been eroded (cut and filled) by a younger layer of matrix-supported cobbles 51 52 and boulders (level OC2). In the north sector, level OC1 is represented by a very thin layer of matrix, 53 54 supported by gravels and cobbles, that disappears towards the northern excavated area, and is overlain by 55 56 a massive fine sandy level (level OC3) and a capping Holocene organic Ap soil (level OC4) (Méndez-Quintas 57 et al. 2019). The post-infrared infrared stimulated luminescence (pIR-IRSL) age (118 ± 9 ka) obtained for 58 59 level OC2 in the south sector provides an estimate for the timing of the debris flow sedimentation and 60 61 localised reworking of the lithic assemblage from an older level or lag deposit (level OC1 or another unknown 62 6 63 64 65 level). The collective chronological evidence available for the Arbo site (sedimentological properties of the Arbo infill sequence, and the pIR-IRSL numerical chronologies) enable us to constrain the original LFA occupation to sometime prior to MIS 5, most likely during MIS 6 and no earlier than MIS7 (Méndez-Quintas 1 2 et al. 2019). 3 4 The lithic industry recovered from both levels OC1 and OC2 are extensive (3142 pieces found during 5 6 excavations and 1826 surface finds) and display similar technological and techno-economic features, which 7 8 is indicative of shared or repeated behavioural trends. The lithic assemblages can be classed as LFA 9 industries due to the use of large flake blanks, and the selected raw materials are quartzite and quartz. 10 11 Quartzite is common in all phases of the chaîne operatoire, while quartz has a complementary role linked to 12 13 flake production. The technological characteristics of the industry are based on some elemental systems of 14 flake production, with a total absence of predetermined schemes associated with the Levallois method. Flake 15 16 tools are of non-standardised types and include some regular shaped LCTs (mainly handaxes and flake 17 18 cleavers), usually finished by the use of soft-hammers (Méndez-Quintas et al. 2019; Méndez Quintas et al. 19 2018) (Fig. 4 a-d). 20 21 22 The Gándaras de Budiño site is the most renowned and cited Galician Palaeolithic site (Aguirre 1964; Aguirre 23 24 and Butzer 1967; Butzer 1967), although its initial interpretation also attracted much scepticism (Méndez- 25 26 Quintas 2007, 2008). The site is located in a +24 m fluvial terrace within the Louro River, which is the 27 westernmost major tributary of the Miño River. The stratigraphic sequence includes several fluvial sand facies 28 29 with extensive accumulations of lithic industry (Méndez-Quintas 2007, 2008; Méndez-Quintas et al. 2018a). 30 31 The lithic industry is made from local (quartz) and non-local (quartzite) raw materials, with quartz being used 32 for flake production, while quartzite was selected to shape more complex artefacts, such as flake tools or 33 34 LCTs. The reduction core patterns are elementary (mainly composed of monopolar or orthogonal cores), 35 36 although there is also an important discoidal component. The flake tools are predominantly of basic types 37 (mainly denticulate and single or transverse scrapers), followed by a large number of LCTs (Fig. 4 e-h), with 38 39 a large number of cleavers and handaxes (Méndez-Quintas 2007, 2008). 40 41 42 The Carvalhas site, in Portugal, is a new site that is currently under investigation, and has yielded large lithic 43 assemblages with LFA features from several loci. The site is located within a fluvial sequence found +21 m 44 45 above the current level of the Miño River, with additional contributions from a lateral alluvial fan (Cunha 46 47 Ribeiro et al. 2017). The artefacts have been knapped on local raw materials, such as quartzite and quartz, 48 which were recovered from fluvial environments. Loci 1 in particular has revealed a large number of knapping 49 50 products associated with a few flake tools and LCTs (Cunha Ribeiro et al. 2017) (Fig. 5 e-i). 51 52 53 The Acheulean site of Porto Maior in Spain similarly documents an extensive accumulation of LCTs (Méndez- 54 Quintas et al. 2018b). The stratigraphic sequence of the site comprises a succession of fluvial layers (labelled 55 56 from PM1 to PM% form bottom to top) belonging to the terrace located +34 m above the current level of the 57 58 Miño River. The chronology of the site has been established by a combination of three numerical dating 59 methods: namely, ESR, pIR-IRSL and single-grain thermally transferred optically stimulated luminescence 60 61 62 7 63 64 65 (TT-OSL) dating. These methods have provided consistent results and constrain the chronology of the main Acheulean levels to between ~300 and 200 ka (Demuro et al. 2019a; Demuro et al. 2019b; Méndez-Quintas et al. 2018b). Bayesian modelling recently further refined this chronology, and indicated that the in situ large 1 2 cutting tool assemblage from level PM4 most likely accumulated during MIS 7 (mean depositional age = 3 210.7 ± 24.7 ka). The “chaînes opératoires” analysis shows that one set of pieces is characterized almost 4 5 exclusively by elements related to the use and discard phase (LCTs and a few flake tools). The number of 6 7 pieces associated with the acquisition and production phases (flakes, waste or cores) is insignificant. The 8 site contains a large number of LCTs (primarily large sized quartzite handaxes; Fig. 5 a-d), without elements 9 10 pertaining to the configuration process (flakes, large blanks or cores). These characteristics suggest that the 11 12 macro-tools were configured elsewhere and were brought to the site for usage and subsequent 13 abandonment. Taphonomic observations indicate that most of the LCTs are found in an autochthonous 14 15 position (Méndez-Quintas et al. 2018b). 16 17 18 Vouga and Lis basins 19 20 21 The central region of the West Iberian study area (Fig. 2) hosts a number of important Acheulean sites 22 (Cunha Ribeiro 1999), including the historical site of Mealhada, which was investigated by C. Ribeiro in 1976 23 24 (Antunes 1988; Cunha Ribeiro 1997). A large mammal faunal assemblage (Homotherium latidens, Elephas 25 26 (Palaeoloxodon) antiquus, Hippopotamus amphibius, Cervus elaphus or Bos primigenius) and some 27 Acheulean artefacts have been described from this site (Antunes 1988). However, subsequent research has 28 29 suggested that the stratigraphic relationship between the mammal fossils and the lithic industry is doubtful, 30 31 hence additional research is needed at this site (Cunha Ribeiro 1997). 32 33 Presently, the area with the most information regarding the Middle Pleistocene archaeological record is the 34 35 Lis basin (Portugal) (Cunha Ribeiro 1999). This basin has 4 main fluvial formations, which range from +72 m 36 37 to +2 m above the current river level, and some colluvium formation overlaying the river terrace sequence 38 (Cunha Ribeiro 1999). The highest fluvial formation (F1, spanning an altitude of +72-50 m) has yielded over 39 40 32 Acheulean assemblages from different localities, the most relevant being the sites of Quinta da Carvalha 41 42 and Outeiro Pelado (Cunha Ribeiro 1999). These assemblages are composed of knapping products and 43 cores, and include a significant number of handaxes or cleavers. The artefact has been made using different 44 45 types of local raw materials, with quartzite being the most readily available rock type. The lithic implements 46 47 display the principal features of the LFA Acheulean, with an elemental reduction system and a large diversity 48 of LCTs, mainly knapped from large flakes supports (Cunha Ribeiro 1999). Several Acheulean sites have 49 50 been excavated in the Lis basin, including Pousias/Quinta do Conego (Fig. 6 a-b), Jardoeira, Casal de Santa 51 52 Maria 1 (Fig. 6 g-h) and Casal do Azemel (Cunha Ribeiro 1999). The first two sites are located in the same 53 fluvial formation, while the other two are found in colluvium environments. The artefacts are made from local 54 55 raw material (mainly quartzite), and include a large number of knapping products and cores associated with 56 57 different percentages of non-standardized flake tools and LCTs, which in turn have a large number of flake 58 supports (Cunha Ribeiro 1999). Casal do Azemel is probably the most important Acheulean site in the region 59 60 because it was excavated over a large area (135 m2) and it has produced 3957 lithic artefacts (Cunha Ribeiro 61 62 8 63 64 65 1999). The raw material used for knapping the artefacts (mainly quartzite) is not local, although it probably came from adjacent valleys in the surrounding area. The site contains a large percentage of LCTs, including 556 handaxes (Fig. 6 c-d) and 127 flake cleavers (Fig. 6 e-f), and it therefore represents the largest LCT 1 2 collection for an excavated site in the Iberian Peninsula. 3 4 Tagus basin 5 6 7 The lower Tagus basin has historically been the main region for studying the Lower Palaeolithic of Portugal 8 9 (Breuil and Zbyszewski 1942; Proença Cunha et al. 2017a; Raposo and Santonja 1995; Zbyszewski 1943, 10 1946). Most of the Acheulean sites are associated with fluvial terraces, although a notable exception is the 11 12 Gruta de Aroeira site (Daura et al. 2017). These sites are mostly associated with the middle regional T4 river 13 14 terrace (+55-26 m), which has been numerically dated to ~340-155 ka (Proença Cunha et al. 2017a). Among 15 the most relevant sites in this basin are the Vale do Forno, in the Alpiarça area (Proença Cunha et al. 2017b; 16 17 Raposo et al. 1985), Ribeira da Ponte da Pedra (Cura 2014), Fonte da Moita (Cura et al. 2013), Gruta de 18 19 Aroeira (Daura et al. 2017; Marks et al. 2002b) and Monte do Famaco (Raposo and Santonja 1995). 20 21 22 The Alpiarça area is the most cited Portuguese locality containing Acheulean sites, because investigations 23 there started in the 1940s (Breuil and Zbyszewski 1942; Zbyszewski 1943, 1946). The sedimentary data 24 25 shows that the large LFA sites are located within and above regional river terrace T4 (Proença Cunha et al. 26 27 2017b). Two main sedimentary beds can be distinguished from published geological data: a basal Lower 28 Gravels Unit (LG) and an overlying Upper Sands Unit (US), as well as additional lateral beds displaying 29 30 paleosoil development (Mozzi 2000; Proença Cunha et al. 2017b). The Vale do Forno contains the sites of 31 32 Vale do Forno 1, 3 (Milharós) and 8 (Proença Cunha et al. 2017b; Raposo 1996). An early study reported 33 Acheulean artefacts in the lower LG unit (Zbyszewski 1946), but the exact stratigraphic position of these finds 34 35 cannot be verified in the present-day setting. Within the upper US unit, the Vale do Forno 1 and 8 sites are 36 37 known, and their age is between ~325 to 200 ka (Proença Cunha et al. 2017b). 38 39 Vale do Forno 1 has provided 254 artefacts made from local quartzite or quartz pebbles and cobbles. The 40 41 assemblage includes flakes, elementary cores, some flake tools and a representative number of handaxes 42 43 and cleavers made from pebble or large flakes (Monteiro Silva Salvador M. M. 2002). The excavated site of 44 Vale do Forno 8 has yielded a large number of artefacts (~3000) within a limited area (20 m2), but the 45 46 assemblage has not been published in detail. The technological features are similar to those from Vale do 47 48 Forno 1, although they come from a higher stratigraphic position within the US unit. The “chaînes opératoires” 49 are complete for flake production and LCTs (Proença Cunha et al. 2017b). The Vale do Forno 3 (Milharós) 50 51 archaeological level is found within a lateral layer situated above the US unit, with an age range of 155-32 ka 52 53 (Proença Cunha et al. 2017b). The lithic assemblage (338 pieces) includes a low percentage of flakes, but a 54 high number of cores, flake tools and LCTs. The cores show low standardized exploitation patterns, although 55 56 some display “proto-levallois” core features. The LCTs, handaxes and cleavers, have a high degree of formal 57 58 standardization and show extensive use of soft-hammer (Raposo 1996; Raposo et al. 1985). This lithic 59 assemblage displays strong similarities with the Arbo site, on the Spanish side of the Miño River basin. 60 61 62 9 63 64 65 The Ribeira da Ponte da Pedra (Portugal) site is located on the T4 terrace and is capped by lateral sediments. The fluvial level is dated to MIS 9-6 and the site has produced 1264 artefacts distributed across 12 lithological units (Cura 2014). The lithic industry is made from local quartzite and displays features that are different to 1 2 those described previously for LFA assemblages in the Tagus region (Cura 2014). The assemblage has a 3 large number of knapping products, elementary cores types, and a few non-standardized flake tools. LCTs 4 5 are present in low percentages (mainly handaxes), though the size of locally available raw material has likely 6 7 limited the production of large flakes. 8 9 The Fonte da Moita site is in a similar geological context to Ribeira da Ponte da Pedra; i.e., within a fluvial 10 11 sequence linked to regional fluvial terrace T4 (Cura et al. 2013). A total of 2852 artefacts made from local 12 13 raw materials (mainly quartzite pebbles and cobbles) have been recovered from this site. The lithic features 14 are very similar to those observed for the Ribeira da Ponte da Pedra assemblage, with elementary core 15 16 reduction patterns, flake tools without standardized morphology, and only a few LCTs (Cura et al. 2013). 17 18 19 Gruta da Aroeira (Portugal) is a particularly relevant site because recent excavations have uncovered a partial 20 21 hominin cranium (Daura et al. 2017), adding to the two dental pieces recovered from earlier excavations 22 (Trinkaus 2003). These hominin remains have a different morphological pattern to other European Middle 23 24 Pleistocene individuals, although some of their features appear to indicate that the specimen represents an 25 26 early (Daura et al. 2017). This site is located within the karst system and belongs to 27 the Almonda archaeological cluster, which also includes the important Palaeolithic site of Gruta da Oliveira 28 29 (Deschamps and Zilhao 2018; Hoffmann et al. 2013). The stratigraphic sequence spans a thickness of 30 31 approximately 4 m, and is divided into 3 major stratigraphic units (Daura et al. 2017; Daura et al. 2018; Marks 32 et al. 2002a; Marks et al. 2002b). The available geochronological data indicate an age range between MIS 33 34 10 and 11 (Daura et al. 2017), while micro and macrofaunal association indicate an advanced Middle 35 36 Pleistocene age (Marks et al. 2002a; Marks et al. 2002b). The large mammals have some anthropic 37 modification (cut-marks) and in some cases preserve evidence of burning. The site therefore possibly 38 39 represents one of the oldest evidences for fire use in the Iberian Peninsula (Daura et al. 2017; Fernández 40 41 Peris et al. 2012; Marks et al. 2002a; Marks et al. 2002b; Roebroeks and Villa 2011). 42 43 The lithic industry comprises 1013 artefacts from the 1998–2002 fieldwork (Marks et al. 2002a; Marks et al. 44 45 2002b), with an additional 393 artefacts recovered from the recent excavations conducted over a 6 m2 area 46 47 (layers X–Xb/c) (Daura et al. 2018). The raw material used includes quartzite and quartz, which are present 48 in similar percentages throughout the studied layers, while there is limited presence of other raw materials 49 50 (e.g., flint). The “chaîne opératoire” shows significant occurrences of knapping products (flakes and debris) 51 52 with few cores, flakes tools and macro-tools (Daura et al. 2018; Marks 2005; Marks et al. 2002a; Marks et al. 53 2002b). The latest studies indicate that the assemblage displays Acheulean features (Daura et al. 2017; 54 55 Daura et al. 2018), but previous investigations have noted significant differences from the main Iberian LFA 56 57 sites (Marks 2005; Marks et al. 2002a; Marks et al. 2002b). We would additionally note that there are some 58 differences between the Gruta da Aroeira lithic implements and other LFA Acheulean artefacts recovered 59 60 from the Western Iberian sites cited here. The presence of more standardized and recurrent shaped flake 61 62 10 63 64 65 tools, with a large degree of retouch (Daura et al. 2018) is especially significant. The reported macro-tools are mainly handaxes (without cleavers or large flakes blanks) displaying intensive knapping, asymmetrical shapes and true biface-tool support. These observations would suggest that the Gruta da Aroeira 1 2 assemblage does not have clear LFA features and, based on present data, we think that it more closely 3 aligns with EMP assemblages seen at a range of other sites of similar age across the Iberian Peninsula (Gran 4 5 Dolina TD 10.1/2, Ambrona middle stratigraphic unit, Cueva de Bolomor or Cuesta de la Bajada). The same 6 7 technological conclusion is applicable to other lithic collections from the Almonda karstic system (Varanda 8 2018). Upstream from Almonda there are several other LFA localities, for example Monte Famaco (Portugal) 9 10 (Raposo and Santonja 1995), but most of these represent sites with surface materials and without clear 11 12 stratigraphy context (Proença Cunha et al. 2017a). 13 14 THE WEST IBERIAN RECORD IN THE CONTEXT OF THE EUROPEAN MIDDLE PLEISTOCENE 15 16 17 The major implication of the synthesis presented here is that the Iberian Atlantic margin preserves a large 18 19 concentration of Middle Pleistocene archaeological sites (Table 1), mainly with LFA features, that collectively 20 21 have regional significance for understanding early human settlement histories of Europe. While early hominin 22 presence has been established for other regions of the Iberian Peninsula (Guadix-Baza basin, Atapuerca, 23 24 for example), similar findings have not been demonstrated unequivocally (with reliable age control) in West 25 26 Iberia until relatively recently. This historical situation did not necessarily reflect the absence of archaeological 27 research in the region, but it has been conditioned by the generally poor conservation of fossil remains at 28 29 many of the open-air sites. The absence of Lower Pleistocene sites in West Iberia is similar to many other 30 31 regions across the European continent, and those Lower Pleistocene sites that have been reported from the 32 western Iberian margins continue to be questioned on archaeological, stratigraphic or chronological grounds 33 34 (Rocca 2016; Roebroeks et al. 2017). 35 36 37 Most of the sites known from this region display clear examples of LFA technology, with the possible 38 exception of Gruta da Aroeira, as detailed above. This situation is very interesting from the perspective of 39 40 understanding European Middle Pleistocene hominin occupation patterns. The broad-scale technological 41 42 affinities between the Acheulean tradition of Southwest Europe and the African Acheulean, especially those 43 identified in the large Atlantic basins of the Iberian Peninsula and Aquitanian area (SW France), have been 44 45 widely documented (Cologne et al. 2013; Santonja and Pérez-González 2010; Santonja and Villa 2006; 46 47 Sharon 2011; Turq et al. 2010). Important differences also exist in the specific technological features of the 48 Acheulean tradition across Europe. In particular, there are notable regional differences in the occurrences of 49 50 cleavers (on flakes), the use of large flake blanks for the configuration of LCTs, and the first occurrence of 51 52 the standardized core reduction pattern (Levallois method) between the Iberian Acheulean tradition and 53 industries found in Northwest Europe (Moncel et al. 2015; Santonja and Pérez-González 2010; Santonja et 54 55 al. 2016; Santonja and Villa 2006; Sharon 2011; Sharon and Barsky 2016). In this context, the site of Porto 56 57 Maior (with its extensive LCT accumulations) provides noteworthy behavioural evidence to support the 58 relationship between the Iberian and African Acheulean industries (Méndez-Quintas et al. 2018b). 59 60 61 62 11 63 64 65 The technological features and age range (~400-150 ka) of the West Iberian Acheulean sites have another important consequence: the chronological coexistence of the LFA Acheulean and EMP in Southwest Europe (Santonja et al. 2016). Over this age range, there are many examples of archaeological sites across the 1 2 region that display distinctly different technological features to those of the African Acheulean tradition; 3 including the middle stratigraphic unit of Ambrona, Level TD10.1-2 at Gran Dolina (Atapuerca), Bolomor 4 5 Cave, Cuesta de la Bajada and possibly Gruta da Aroeira (Ollé et al. 2013; Santonja et al. 2014; Santonja 6 7 et al. 2016). Among the main technological features found at these sites are systematically and pre- 8 configured core exploitation patterns (as Levallois or Quina types), “chaîne opératoire” ramifications, 9 10 repetitive and standardized flake tools, and limited occurrences (or total absences) of macro-tools (mainly 11 12 recovered as handaxe-tool support) (Ashton and Scott 2015; Hérisson et al. 2016b; Locht et al. 2016; 13 Santonja et al. 2016; Turq et al. 2010). 14 15 16 These cultural patterns have important implications for understanding the complex human occupation history 17 18 of the continent. It is increasingly clear from the paleoanthropological record that a mosaic of human 19 populations of different evolutionary relations and geographic origins, with varied biological and technological 20 21 attributions, likely co-existed in Europe at various times during the Middle Pleistocene (Bermúdez de Castro 22 23 and Martinón-Torres 2013; Bermúdez de Castro et al. 2018; Daura et al. 2017; Dennell et al. 2011; Hublin 24 2009; Martinón-Torres et al. 2018; Vialet et al. 2018). The temporal and geographical coexistence of LFA 25 26 Acheulean and EMP sites in the region suggests the coexistence of culturally different human populations of 27 28 varied geographical origins. In this sense, the African affinities of the Acheulean sites may be consistent with 29 30 a technology brought in by an Acheulean "intrusive" population, against a backdrop of “native” hominin groups 31 that evolved within Europe during the Lower or ancient Middle Pleistocene. 32 33 34 CONCLUSION 35 36 37 This paper has highlighted how the Iberian Atlantic margin is a key area for the study of hominin population 38 dynamics during the European Middle Pleistocene. This region has an abundant archaeological record, with 39 40 a strong geochronological foundation spanning MIS 9-6. The cultural and technological features of the region 41 42 are consistent with broader patterns across Southwest Europe. The African affinities of Acheulean sites 43 across the region re-invigorate old debates about the age and geographical origin of the southwestern 44 45 European Acheulean. The observed technological diversity across the broader Iberian Peninsula during the 46 47 Middle Pleistocene mirrors the region’s complex paleoanthropological record, and suggests the possible 48 coexistence of distinctive hominin groups with different technologies and geographic origins. 49 50 51 ACKNOWLEDGEMENTS 52 53 54 E.M-Q. is funded by Post-Doc Xunta de Galicia Grant ED481B-2018/063. M. Demuro. is funded by Australian 55 56 Research Council (ARC) Discovery Early Career Researcher Award DE160100743 and L.J.A. is funded by 57 ARC Future Fellowship project FT130100195. M. Duval is funded by ARC Future Fellowship Grant 58 59 FT150100215. All authors collaborate within the framework of the project Miño/Minho, Os primeiros 60 61 habitantes do Baixo Minho. Estudo das ocupações pleistocénicas da região. The authors appreciate the work 62 12 63 64 65 of Raquel Rojas Mendoza in the drawing of lithic industry of the As Carvalhas site. All the authors thank the UISPP session organizers, Roxane Rocca, Jordi Serangeli, Vincent Lhomme, for the invitation to participate in this special issue and we thank the anonymous reviewers for their constructive comments and suggestions, 1 2 which helped to improve this paper. 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 13 63 64 65 DECLARATION A CONFLICT OF INTEREST

1 On behalf of all authors, the corresponding author states that there is no conflict of interest. 2 3 4 5 6 7 8 9

10 11 12 13 14 15 Eduardo Méndez-Quintas 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 14 63 64 65 REFERENCES

Aguirre, E. (1964). Las Gándaras de Budiño Porriño (Pontevedra) (Vol. 31). Madrid: Excavaciones Arqueologicas en España 31. 1 2 3 Aguirre, E., & Butzer, K. W. (1967). Problematical Pleistocene artifact assemblage from Northwestern Spain. 4 Science, 157(3787), 430-431. 5 6 Álvarez-Alonso, D. (2014). First Neanderthal settlements in northern Iberia: The Acheulean and the 7 8 emergence of Mousterian technology in the Cantabrian region. Quaternary International, 326-327, 9 288-306, doi:10.1016/j.quaint.2012.12.023. 10 11 Álvarez Blázquez, X. M., & Bouza Brey, F. (1949). Industrias Paleolíticas de la comarca de Tuy. Cuadernos 12 13 de Estudios Gallegos, XVI, 5-42. 14 15 Ameijenda Iglesias, A., Lombera Hermida, A., Pérez Alberti, A., Rodríguez-Álvarez, X. P., & Fábregas 16 Valcarce, R. (2010). Geomorphological and Geoarchaeological evolution of the Monforte de Lemos 17 Basin (Galicia, Spain). Erosion phases and post-depositional processes in NW Iberia. Estudos do 18 19 Quaternario, 6, 5-22. 20 21 Antoine, P., Moncel, M. H., Voinchet, P., Locht, J. L., Amselem, D., Herisson, D., et al. (2019). The earliest 22 evidence of Acheulian occupation in Northwest Europe and the rediscovery of the Moulin Quignon 23 24 site, Somme valley, France. Scientific Reports, 9(1), 13091, doi:10.1038/s41598-019-49400-w. 25 26 Antunes, M. A. (1988). Présence de Hippopotamus incognitus au Portugal et remarques sur les Sites 27 Quaternaires de Mealhada. Comunicações dos servicios geologicos de Portugal, 74, 165-172. 28 29 30 Arsuaga, J. L., Martinez, I., Arnold, L. J., Aranburu, A., Gracia-Tellez, A., Sharp, W. D., et al. (2014). 31 Neandertal roots: Cranial and chronological evidence from Sima de los Huesos. Science, 344(6190), 32 1358-1363, doi:10.1126/science.1253958. 33 34 35 Ashton, N., & Scott, B. (2015). The British Middle Palaeolithic. Quaternary International, 36 doi:10.1016/j.quaint.2015.06.011. 37 38 39 Bermúdez de Castro, J. M., & Martinón-Torres, M. (2013). A new model for the evolution of the human 40 Pleistocene populations of Europe. Quaternary International, 295, 102-112, 41 doi:10.1016/j.quaint.2012.02.036. 42 43 44 Bermúdez de Castro, J. M., Martinón-Torres, M., Martínez de Pinillos, M., García-Campos, C., Modesto- 45 Mata, M., Martín-Francés, L., et al. (2018). Metric and morphological comparison between the Arago 46 (France) and Atapuerca-Sima de los Huesos (Spain) dental samples, and the origin of . 47 Quaternary Science Reviews, doi:10.1016/j.quascirev.2018.04.003. 48 49 50 Bermúdez de Castro, J. M., Martinón TorresMartinón-Torres, M., Rosell, J., Blasco, R., Arsuaga, J. L., & 51 Carbonell, E. (2016). Continuity versus discontinuity of the human settlement of Europe between the 52 late Early Pleistocene and the early Middle Pleistocene. The mandibular evidence. Quaternary 53 Science Reviews, 153, 51-62, doi:10.1016/j.quascirev.2016.10.010. 54 55 56 Bouza Brey, F., & Álvarez Blázquez, X. M. (1954). Industrias paleolíticas do Baixo Miño (concellos de A 57 Guarda, O Rosal, Tomiño e Oia). Trabalhos de Antropologia e Etnologia, XIV(1-4), 5-65. 58 59 60 61 62 15 63 64 65 Breuil, A. H., & Zbyszewski, G. (1942). Contribution à l’étude des industries paléolithiques du Portugal et leurs rapports avec la géologie du Quaternaire. Les principaux gisements des deux rives de l’ancien estuaire du Tage. Comunicações dos servicios geologicos de Portugal, XXIII(I), 1-374.

1 2 Butzer, K. W. (1967). Geomorphology and Stratigraphy of the Palaeolithic site of Budiño. Eiszeitalter und 3 Gegenwart, 18(31-32), 82-103. 4 5 Cano Pan, J. A., Aguirre, E., Giles Pacheco, F., Gracia Prieto, F. J., Santiago Pérez, A., Mata Almonte, E., 6 et al. (1999). Secuencia cronoestratigráfica y poblamiento paleolítico en la cuenca Media-Baja de Río 7 8 Miño. Sector Tuy (Pontevedra)-Cortegada (Ourense). In L. Pallí Buxó, & C. Roqué i Pau (Eds.), 9 Avances en el estudio del Cuaternario español (secuencias, indicadores paleambientales y evolución 10 de procesos) (pp. 231-240). 11 12 13 Cano Pan, J. A., Aguirre, E., Giles Pacheco, F., Gracia Prieto, J., Santiago Pérez, A., Mata Almonte, E., et 14 al. (1997). Evolución del Pleistoceno en la cuenca baja del Miño, sector La Guardia-Tuy. Secuencia 15 de los primeros poblamientos humanos y registro arqueológico. In J. Rodríguez Vidal (Ed.), 16 Cuaternario Ibérico (pp. 201-212). 17 18 19 Carbonell, E., Bermúdez de Castro, J. M., Arsuaga, J., Díez Fernández-Lomana, J. C., Rosas, A., Cuenca- 20 Bescós, G., et al. (1995). Lower Pleistocene hominids and artifacts from Atapuerca-TD6 (Spain). 21 Science, 269, 826-830, doi:10.1126/science.7638598. 22 23 24 Carbonell, E., Bermúdez de Castro, J. M., Parés, J. M., Pérez-González, A., Cuenca-Bescós, G., Ollé, A., et 25 al. (2008). The first hominin of Europe. Nature, 452(7186), 465-469, doi:10.1038/nature06815. 26 27 Cardoso, J. L. (1993). Contribuição para o conhecemento dos grandes mamiferos do Plistocenico Superior 28 de Portugal: Cámara Municipal de Oeiras. 29 30 31 Chauhan, P. R., Bridgland, D. R., Moncel, M.-H., Antoine, P., Bahain, J.-J., Briant, R., et al. (2017). Fluvial 32 deposits as an archive of early human activity: Progress during the 20 years of the Fluvial Archives 33 Group. Quaternary Science Reviews, doi:10.1016/j.quascirev.2017.03.016. 34 35 36 Cologne, D., Hernádez, M., Lelouvier, L.-A., Mercier, N., Mourre, V., & Busseuil, N. (2013). Paléolithique 37 ancien et Paléolithique moyen ancien dans le piémont pyrénéen occidental: relations entre chaînes 38 opératoires, continuités et ruptures. In J. Jaubert, N. Forument, & P. Depaepe (Eds.), Transitions, 39 ruptures et continuité en Préhistoire. Volume 1 (pp. 119-137): Société préhistorique française. 40 41 42 Cunha, P. P., Martins, A. A., Huot, S., Murray, A., & Raposo, L. (2008). Dating the Tejo river lower terraces 43 in the Ródão area (Portugal) to assess the role of tectonics and uplift. Geomorphology, 102(1), 43- 44 54. 45 46 47 Cunha Ribeiro, J.-P. (1997). A estação paleolítica da Mealhada nos 120 anos de estudo do Acheulense em 48 Portugal. O Arqueólogo Portugues, Serie IV. 13/15, 35-52. 49 50 51 Cunha Ribeiro, J.-P. (1999). O acheulense no centro de Portugal: o Vale do Lis. Contribuicão para uma 52 abordagem tecno-tipológica das suas indústrias líticas e problemática do seu contexto 53 cronoestratigráfico. Universidade de Lisboa, Lisboa. 54 55 56 Cunha Ribeiro, J.-P., Monteiro-Rodrigues, S., Gomes, A., Méndez Quintas, E., Meireles, J., Pérez-González, 57 A., et al. (2017). Ocupações Pleistocénicas na margem esquerda do Baixo Minho (Miño/Minho 2). 58 Objetivos e primeiros resultados de um projeto transfronteiriço. In J. M. Arnaud, & A. Martins (Eds.), 59 Arqueologia em Portugal / 2017-Estado da Questão (pp. 303-318): Associação dos Arqueólogos 60 Portugueses. 61 62 16 63 64 65 Cura, S. (2014). Tecnologia lítica e comportamento humano no pleistocénico médio final do Alto Ribatejo: estudo da indústria lítica da Ribeira da Ponte da Pedra. Universidade de Trás-os-Montes e Alto Douro,

1 2 Cura, S., Rosina, P., Grimaldi, S., & Osterbeek, L. (2013). Os sítios da Ribeira da Ponte da Pedra e da Fonte 3 da Moita no contexto das mais antigas ocupações do vale do Tejo. Arkeos, 34, 33-44. 4 5 da Silva, P. X. (2012). A jazida Paleolítica de Marinho (Afife): um ensaio de aplicação do sistema lógico- 6 analítico às indústrias paleolíticas do NO Peninsular. Universidade do Minho, 7 8 9 Daura, J., Sanz, M., Arsuaga, J. L., Hoffmann, D. L., Quam, R. M., Ortega, M. C., et al. (2017). New Middle 10 Pleistocene hominin cranium from Gruta da Aroeira (Portugal). Proceedings of the National Academy 11 of Sciences, 201619040, doi:10.1073/pnas.1619040114. 12 13 14 Daura, J., Sanz, M., Deschamps, M., Matias, H., Igreja, M., Villaescusa, L., et al. (2018). A 400,000-year-old 15 Acheulean assemblage associated with the Aroeira-3 human cranium (Gruta da Aroeira, Almonda 16 karst system, Portugal). Comptes Rendus Palevol, doi:10.1016/j.crpv.2018.03.003. 17 18 19 Demuro, M., Arnold, L., Duval, M., Spooner, N. A., Creighton, D., Méndez-Quintas, E., et al. (2020). Insights 20 into the relationship between luminescence and ESR dating signals from Spanish sedimentary quartz 21 samples of different geologic origins. Quaternary International. 22 23 24 Demuro, M., Arnold, L. J., Duval, M., Méndez-Quintas, E., Santonja, M., & Pérez-González, A. (2020). 25 Refining the chronology of Acheulean deposits at Porto Maior in the River Miño basin (Galicia, Spain) 26 using a comparative luminescence and ESR dating approach. Quaternary International. 27 https://doi.org/10.1016/j.quaint.2020.01.005. 28 29 30 Dennell, R. W., Martinón-Torres, M., & Bermúdez de Castro, J. M. (2011). Hominin variability, climatic 31 instability and population demography in Middle Pleistocene Europe. Quaternary Science Reviews, 32 30(11-12), 1511-1524, doi:10.1016/j.quascirev.2009.11.027. 33 34 35 Deschamps, M., & Zilhao, J. (2018). Assessing site formation and assemblage integrity through 36 refitting at Gruta da Oliveira (Almonda karst system, , Portugal): A Middle 37 case study. PLOS ONE, 13(2), e0192423, doi:10.1371/journal.pone.0192423. 38 39 40 Falguères, C., Bahain, J.-J., Pérez-González, A., Mercier, N., Santonja, M., & Dolo, J.-M. (2006). The Lower 41 Acheulian site of Ambrona, Soria (Spain): ages derived from a combined ESR/U-series model. 42 Journal of Archaeological Science, 33(2), 149-157, doi:10.1016/j.jas.2005.07.006. 43 44 45 Fernández Peris, J., González, V. B., Blasco, R., Cuartero, F., Fluck, H., Sañudo, P., et al. (2012). The 46 earliest evidence of hearths in Southern Europe: The case of Bolomor Cave (Valencia, Spain). 47 Quaternary International, 247, 267-277, doi:10.1016/j.quaint.2010.10.014. 48 49 50 Harvati, K., Roding, C., Bosman, A. M., Karakostis, F. A., Grun, R., Stringer, C., et al. (2019). Apidima Cave 51 fossils provide earliest evidence of Homo sapiens in Eurasia. Nature, doi:10.1038/s41586-019-1376- 52 z. 53 54 55 Hérisson, D., Brenet, M., Cliquet, D., Moncel, M.-H., Richter, J., Scott, B., et al. (2016a). The emergence of 56 the Middle Palaeolithic in north-western Europe and its southern fringes. Quaternary International, 57 411, 233-283, doi:10.1016/j.quaint.2016.02.049. 58 59 60 61 62 17 63 64 65 Hérisson, D., Coutard, S., Goval, E., Locht, J. L., Antoine, P., Chantreau, Y., et al. (2016b). A new key-site for the end of Lower Palaeolithic and the onset of Middle Palaeolithic at Etricourt-Manancourt (Somme, France). Quaternary International, 409(B), 73-91, doi:10.1016/j.quaint.2016.01.055.

1 2 Hoffmann, D. L., Pike, A. W. G., Wainer, K., & Zilhão, J. (2013). New U-series results for the speleogenesis 3 and the Palaeolithic archaeology of the Almonda karstic system (Torres Novas, Portugal). Quaternary 4 International, 294, 168–182, doi:10.1016/j.quaint.2012.05.027. 5 6 Hublin, J. J. (2009). The origin of Neandertals. Proceedings of the National Academy of Sciences, 106(38), 7 8 16022-16027, doi:10.1073/pnas.0904119106. 9 10 Lisiecki, L.E., Raymo, M.E., (2005). A Pliocene–Pleistocene stack of 57 globally distributed benthic d18O 11 records. Paleoceanography 20, PA1003. 12 13 14 Locht, J.-L., Hérisson, D., Goval, E., Cliquet, D., Huet, B., Coutard, S., et al. (2016). Timescales, space and 15 culture during the Middle Palaeolithic in northwestern France. Quaternary International, 411, 129-148, 16 doi:http://dx.doi.org/10.1016/j.quaint.2015.07.053. 17 18 19 Marks, A. E. (2005). Micoquian elements in the Portuguese middle pleistocene: assemblages from the galeria 20 pesada. In N. Ferreira Bicho (Ed.), O Paleolítico: actas do IV Congresso de Arqueologia Peninsular 21 (Faro, 14 a 19 de Setembro de 2004) (pp. 195-206). 22 23 24 Marks, A. E., Brugal, J.-P., Chabai, V. P., Monigal, K., Goldberg, P., Hockett, H., et al. (2002a). Le gisement 25 pléistocène moyen de Galeria Pesada (Estrémadure, Portugal) : premiers résultats. Paleo, 14, 77- 26 100. 27 28 29 Marks, A. E., Monigal, K., Chabai, V. P., Brugal, J., Goldberg, P., Hockett, H., et al. (2002b). Excavations at 30 the Middle Pleistocene Cave Site of Galeria Pesada, Portuguese Estremadura: 1997–1999. O 31 Arqueólogo Portugues, Serie IV(20), 7-38. 32 33 34 Martínez, K., Garcia, J., Carbonell, E., Agustí, J., Bahain, J.-J., Blain, H.-A., et al. (2010). A new Lower 35 Pleistocene archeological site in Europe (Vallparadís, Barcelona, Spain). Proceedings of the National 36 Academy of Sciences, 107(13), 5762-5767, doi:10.1073/pnas.0913856107. 37 38 39 Martinón-Torres, M., Xing, S., Liu, W., & Bermúdez de Castro, J. M. (2018). A “source and sink” model for 40 East Asia? Preliminary approach through the dental evidence. Comptes Rendus Palevol, 17(1-2), 33- 41 43, doi:10.1016/j.crpv.2015.09.011. 42 43 44 Martins, A. A., Cunha, P. P., Buylaert, J.-P., Huot, S., Murray, A. S., Dinis, P., et al. (2010). K-Feldspar IRSL 45 dating of a Pleistocene river terrace staircase sequence of the Lower Tejo River (Portugal, western 46 Iberia). Quaternary Geochronology, 5(2-3), 176-180. 47 48 49 Meireles, J. (1992). As industrias liticas do litoral minhoto. Contexto cronoestratigrafico e paleoambiental. 50 Cadernos de Arqueologia, Monografias 7, 577. 51 52 53 Méndez-Quintas, E. (2007). El yacimiento achelense de As Gándaras de Budiño. La industria en facies 54 fluviales. Complutum, 18, 27-45. 55 56 Méndez-Quintas, E. (2008). La industria lítica de las facies coluviales del yacimiento achelense de As 57 Gándaras de Budiño. El Locus V en las excavaciones de Vidal Encinas. Zephyrus, LXIII(2), 41-61. 58 59 60 61 62 18 63 64 65 Méndez-Quintas, E., Demuro, M., Arnold, L. J., Duval, M., Pérez-González, A., & Santonja, M. (2019). Insights into the late stages of the Acheulean technocomplex of Western Iberia from the Arbo site (Galicia, Spain). Journal of Archaeological Science: Reports, 27, 101934, doi:10.1016/j.jasrep.2019.101934. 1 2 3 Méndez-Quintas, E., Santonja , M., Pérez-González, A., Arnold, L., Cunha Ribeiro, J.-P., Demuro, M., et al. 4 (2018a). El yacimiento achelense de As Gándaras de Budiño: Síntesis y perspectivas después de 50 5 años de desencuentros. Estudos do Quaternario, 19, 1-22, doi:10.30893/eq.v0i19.182. 6 7 8 Méndez-Quintas, E., Santonja, M., Perez-Gonzalez, A., Duval, M., Demuro, M., & Arnold, L. J. (2018b). First 9 evidence of an extensive Acheulean large cutting tool accumulation in Europe from Porto Maior 10 (Galicia, Spain). Scientific Reports, 8(1), 3082, doi:10.1038/s41598-018-21320-1. 11 12 13 Méndez Quintas, E., Pérez-González, A., Ledo Bernárdez, M., & Serodio Domínguez, A. (2018). La industria 14 lítica del yacimiento achelense de Arbo (Pontevedra). Variables del paleolítico antiguo de Galicia en 15 el contexto peninsular. Zephyrus, LXXXII, 3-34, doi:10.14201/zephyrus201882334. 16 17 Méndez Quintas, E., Santonja, M., Pérez-González, A., Demuro, M., Duval, M., & Arnold, L. (2017). O 18 19 sistema de terraços da bacia baixa do rio Minho (margem galega). Paper presented at the 8º 20 Congresso Nacional de Geomorfología. Geomorfología 2017, Facultade de Letras de Porto, 21 22 Moncel, M.-H., Arzarello, M., Boëda, É., Bonilauri, S., Chevrier, B., Gaillard, C., et al. (2018). The 23 24 assemblages with bifacial tools in Eurasia (first part). What is going on in the West? Data on western 25 and southern Europe and the Levant. Comptes Rendus Palevol, 17(1-2), 45-60, 26 doi:10.1016/j.crpv.2015.09.009. 27 28 Moncel, M.-H., Ashton, N., Lamotte, A., Tuffreau, A., Cliquet, D., & Despriée, J. (2015). The Early Acheulian 29 30 of north-western Europe. Journal of Anthropological Archaeology, 40, 302-331, 31 doi:10.1016/j.jaa.2015.09.005. 32 33 Monteiro-Rodrigues, S., & Cunha Ribeiro, J.-P. (2014). A Estação Paleolítica do Cerro (Vila Nova de Gaia, 34 35 Noroeste de Portugal): caracterização preliminar dos utensílios com configuração bifacial. Estudos 36 do Quaternario, 11, 3-18. 37 38 Monteiro-Rodrigues, S., Gomes, A., & Cunha Ribeiro, J.-P. (2016). Os artefactos líticos talhados do litoral 39 de Vila Nova de Gaia: caracterização genérica, contextualização geoarqueológica, problemas e 40 41 perspectivas. Estudos do Quaternario, 14, 99-116. 42 43 Monteiro-Rodrigues, S., & González, A. (2010). A estação paleolítica da Praia da Aguda (Arcozelo, Vila Nova 44 de Gaia). Notícia preliminar. Estudos do Quaternario, 6, 23-36. 45 46 47 Monteiro Silva Salvador M. M. (2002). Contribução para o estudo do Paleolitico Inferior do Vale do Forno- 48 Alpiarça, no seu contexto crono-estratigrafico. Universidade de Lisboa, 49 50 51 Moreno, D., Falguères, C., Pérez-González, A., Duval, M., Voinchet, P., Benito-Calvo, A., et al. (2012). ESR 52 chronology of alluvial deposits in the Arlanzón valley (Atapuerca, Spain): Contemporaneity with 53 Atapuerca Gran Dolina site. Quaternary Geochronology, 10, 418-423, 54 doi:10.1016/j.quageo.2012.04.018. 55 56 57 Mosquera, M., Ollé, A., Rodríguez-Álvarez, X. P., & Carbonell, E. (2018). Shedding light on the Early 58 Pleistocene of TD6 (Gran Dolina, Atapuerca, Spain): The technological sequence and occupational 59 inferences. PLOS ONE, 13(1), e0190889, doi:10.1371/journal.pone.0190889. 60 61 62 19 63 64 65 Mozzi, P. (2000). Middle Terrace Deposits of the Tagus River in Alpiarça, Portugal, in Relation to Early Human Occupation. Quaternary Research, 54(3), 359-371, doi:10.1006/qres.2000.2154.

Muttoni, G., Scardia, G., & Kent, D. V. (2018). Early hominins in Europe: The Galerian migration hypothesis. 1 2 Quaternary Science Reviews, 180, 1-29, doi:10.1016/j.quascirev.2017.10.031. 3 4 Ollé, A., Mosquera, M., Rodríguez, X. P., de Lombera-Hermida, A., García-Antón, M. D., García-Medrano, 5 P., et al. (2013). The Early and Middle Pleistocene technological record from Sierra de Atapuerca 6 (Burgos, Spain). Quaternary International, 295, 138-167, doi:10.1016/j.quaint.2011.11.009. 7 8 9 Pereira, A., Nomade, S., Moncel, M.-H., Voinchet, P., Bahain, J.-J., Biddittu, I., et al. (2018). Integrated 10 geochronology of Acheulian sites from the southern Latium (central Italy): Insights on human- 11 environment interaction and the technological innovations during the MIS 11-MIS 10 period. 12 13 Quaternary Science Reviews, 187, 112-129, doi:10.1016/j.quascirev.2018.03.021. 14 15 Proença Cunha, P., Cura, S., Cunha Ribeiro, J.-P., Figueiredo, S., Martins, A., Raposo, L., et al. (2017a). As 16 indústrias do Paleolítico Inferior e Médio associadas ao Terraço T4 do Baixo Tejo (Portugal central); 17 Arquivos da mais antiga ocupação humana no oeste da Ibéria, com ca. 340 ka a 155 ka. Journal of 18 19 Lithic Studies, 4(3), 27-56, doi:10.2218/jls.v4i3.2531. 20 21 Proença Cunha, P., Martins, A. A., Buylaert, J.-P., Murray, A. S., Raposo, L., Mozzi, P., et al. (2017b). New 22 data on the chronology of the Vale do Forno sedimentary sequence (Lower Tejo River terrace 23 24 staircase) and its relevance as a fluvial archive of the Middle Pleistocene in western Iberia. Quaternary 25 Science Reviews, 166, 204-226, doi:10.1016/j.quascirev.2016.11.001. 26 27 Ramil Rego, E., & Ramil Soneira, J. (Eds.). (2014). Louselas (Ribadeo, Lugo). Los inicios del poblamiento 28 humano en la región cantábrica (Vol. 6, Monografías MUPAV 6). 29 30 31 Ramos, A. M., Cunha, P. P., Cunha, L. S., Gomes, A., Lopes, F. C., Buylaert, J.-P., et al. (2012). The River 32 Mondego terraces at the Figueira da Foz coastal area (western central Portugal): Geomorphological 33 and sedimentological characterization of a terrace staircase affected by differential uplift and glacio- 34 35 eustasy. Geomorphology, 165-166, 107-123, doi:10.1016/j.geomorph.2012.03.037. 36 37 Raposo, L. (1985). Le paléolithique inférieur archaïque au Portugal. Bulletin de la Société Préhistorique 38 Française, 173-180. 39 40 41 Raposo, L. (1996). Quartzite bifaces and cleaver in the final acheulian assembalge of Milharós (Alpiarça, 42 Portugal). In N. Moloney, L. Raposo, & M. Santonja (Eds.), Non-Flint Stone Tools and the Palaeolithic 43 Occupation of the Iberian Peninsula (pp. 151-165, BAR International Series 649). 44 45 46 Raposo, L., & Cardoso, J. L. (2000). A questão das primeiras ocupações humanas do territorio portugues, 47 no quadro europeu e circum-mediterranico: história das investigações, situação actual, perspectivas 48 futuras. Estudos do Quaternario, 3, 57-72. 49 50 51 Raposo, L., Carreira, J. R., & Salvador, M. (1985). A estação acheulense de Milharõs, Val de Forno, Alpiarça. 52 Paper presented at the I Reunião do Quaternário Ibérico, Lisboa 1985, 53 54 55 Raposo, L., & Carrera, J. R. (1986). Acerca da existência de complexos industriais pré-acheulenses no 56 território português. O Arqueólogo Portugues, 4, 7-90. 57 58 59 Raposo, L., & Santonja, M. (1995). The earliest occupation of Europe: the Iberian peninsula. In W. 60 Roebroeks, & T. V. Kolfschoten (Eds.), The earliest occupation of Europe (pp. 7-25): University of 61 Leiden. 62 20 63 64 65 Rocca, R. (2016). First settlements in Central Europe: Between originality and banality. Quaternary International, 409(B), 213-221, doi:10.1016/j.quaint.2015.08.066.

Rodríguez Asensio, J. A. (Ed.). (1983). La presencia humana más antigua en Asturias: (el Paleolítico Inferior 1 2 y Medio) (Estudios de Arqueología Asturiana 2). Oviedo. 3 4 Rodríguez Gracia, V. (1976). Notas sobre el yacimiento paleolítico de A Piteira, Toén (Orense). Boletín 5 Auriense, VI, 25-42. 6 7 8 Roebroeks, W., Gaudzinski-Windheuser, S., Baales, M., & Kahlke, R.-D. (2017). Uneven Data Quality and 9 the Earliest Occupation of Europe—the Case of Untermassfeld (Germany). Journal of Paleolithic 10 Archaeology, doi:10.1007/s41982-017-0003-5. 11 12 13 Roebroeks, W., & Villa, P. (2011). On the earliest evidence for habitual use of fire in Europe. Proceedings of 14 the National Academy of Sciences, 108(13), 5209-5214. 15 16 17 Rosina, P., Voinchet, P., Bahain, J.-J., Cristovão, J., & Falgueres, C. (2014). Dating the onset of Lower Tagus 18 River terrace formation using electron spin resonance. Journal of Quaternary Science, 29(2), 153- 19 162, doi:10.1002/jqs.2689. 20 21 22 Rubio-Jara, S., Panera, J., Rodríguez-de-Tembleque, J., Santonja, M., & Pérez-González, A. (2016). Large 23 flake Acheulean in the middle of Tagus basin (Spain): Middle stretch of the river Tagus valley and 24 lower stretches of the rivers Jarama and Manzanares valleys. Quaternary International, 411, 349- 25 366, doi:10.1016/j.quaint.2015.12.023. 26 27 28 Sánchez-Cervera, B., Santonja Gómez, M., Pérez-González, A., Domínguez-Rodrigo, M., & Sánchez- 29 Romero, L. (2015). La industria lítica del yacimiento achelense de Torralba (Soria, España). 30 Colecciones marqués de Cerralbo y Howell. Trabajos de Prehistoria, 72(1), 41-63, 31 doi:10.3989/tp.2015.12143. 32 33 34 Santonja, M., López Martínez, L., & Pérez-González, A. (Eds.). (1980). Ocupaciones achelenses en el Valle 35 del Jarama (Arganda, Madrid) (Arqueología y Paleoecología, Vol. 1): Diputación Provincial de Madrid. 36 37 38 Santonja, M., & Pérez-González, A. (2010). Mid-Pleistocene Acheulean industrial complex in the Iberian 39 Peninsula. Quaternary International, 223-224, 154-161, doi:j.quaint.2010.02.010. 40 41 42 Santonja , M., Pérez-González, A., Dominguez- Rodrigo, M., Panera Gallego, J., Rubio Jara, S., Sesé, C., 43 et al. (2014). The Middle Paleolithic site of Cuesta de la Bajada (Teruel, Spain): a perspective on the 44 Acheulean and Middle Paleolithic technocomplexes in Europe. Journal of Archaeological Science, 45 49, 556-571, doi:10.1016/j.jas.2014.06.003. 46 47 48 Santonja, M., Pérez-González, A., Panera, J., Rubio-Jara, S., & Méndez-Quintas, E. (2016). The coexistence 49 of Acheulean and Ancient Middle Palaeolithic techno-complexes in the Middle Pleistocene of the 50 Iberian Peninsula. Quaternary International, 411, 367-377, doi:10.1016/j.quaint.2015.04.056. 51 52 53 Santonja, M., Rubio Jara, S., Panera, J., Sánchez-Romero, L., Tarriño, A., & Pérez-González, A. (2017). 54 Ambrona revisited: The Acheulean lithic industry of the Lower Stratigraphic Complex. Quaternary 55 International, 480, 95-117, doi:10.1016/j.quaint.2017.01.021. 56 57 Santonja, M., & Villa, P. (2006). The Acheulean in Southwestern Europe. In N. Goren-Inbar, & G. Sharon 58 59 (Eds.), Axe Age. Acheulian Tool-making from Quarry to Discard (pp. 429-478): Equinox Publishers. 60 61 62 21 63 64 65 Sharon, G. (2010). Large flake Acheulian. Quaternary International, 223-224, 226-233, doi:10.1016/j.quaint.2009.11.023.

Sharon, G. (2011). Flakes Crossing the Straits? Entame Flakes and Northern Africa–Iberia contact during 1 2 the Acheulean. African Archaeological Review, 1-16, doi:10.1007/s10437-011-9087-3. 3 4 Sharon, G., & Barsky, D. (2016). The emergence of the Acheulian in Europe – A look from the east. 5 Quaternary International, 411, 25-33, doi:10.1016/j.quaint.2015.11.108. 6 7 8 Silva, P. G., Roquero, E., López-Recio, M., Huerta, P., & Martínez-Graña, A. M. (2017). Chronology of fluvial 9 terrace sequences for large Atlantic rivers in the Iberian Peninsula (Upper Tagus and Duero drainage 10 basins, Central Spain). Quaternary Science Reviews, 166, 188-203, 11 doi:10.1016/j.quascirev.2016.05.027. 12 13 14 Soriano, S., & Villa, P. (2017). Early Levallois and the beginning of the Middle Paleolithic in central Italy. 15 PLOS ONE, 12(10), e0186082, doi:10.1371/journal.pone.0186082. 16 17 18 Sousa, M. F., & Figueiredo, S. M. (2001). The Pleistocene elephants of Portugal. In G. Cavarreta, P. Gioia, 19 M. Mussi, & M. R. Palombo (Eds.), The World of Elephants: Proceedings of the 1st International 20 Congress (pp. 611-615). Roma: Consiglio Nazionale delle Ricerche. 21 22 23 Texier, J.-P., Cunha Ribeiro, J.-P., & Meireles, J. (1995). L'Acheuléen du nord et du centre du Portugal: bilan 24 des connaissances actuelles. Paleo. supplément, 1(1), 185-193. 25 26 27 Toro-Moyano, I., Barsky, D., Cauche, D., Celiberti, V., Grégoire, S., Lebegue, F., et al. (2011). The archaic 28 stone tool industry from Barranco León and Fuente Nueva 3, (Orce, Spain): Evidence of the earliest 29 hominin presence in southern Europe. Quaternary International, 243(1), 80-91, 30 doi:10.1016/j.quaint.2010.12.011. 31 32 33 Toro-Moyano, I., Martínez-Navarro, B., Agustí, J., Souday, C., Bermúdez de Castro, J. M., Martinón-Torres, 34 M., et al. (2013). The oldest human fossil in Europe dated to ca. 1.4 Ma at Orce (Spain). Journal of 35 Human Evolution, 65(1), 1-9, doi:10.1016/j.jhevol.2013.01.012. 36 37 38 Trinkaus, E. (2003). Later Middle Pleistocene human remains from the Almonda Karstic system, Torres 39 Novas, Portugal. Journal of Human Evolution, 45(3), 219-226, doi:10.1016/j.jhevol.2003.07.001. 40 41 42 Turq, A., Brenet, M., Colonge, D., Jarry, M., Lelouvier, L.-A., O'Farrell, M., et al. (2010). The first human 43 occupations in southwestern France: A revised summary twenty years after the Abbeville/Saint 44 Riquier colloquium. Quaternary International, 223-224, 383-398. 45 46 Vallverdu, J., Saladie, P., Rosas, A., Huguet, R., Caceres, I., Mosquera, M., et al. (2014). Age and date for 47 48 early arrival of the Acheulian in Europe (Barranc de la Boella, la Canonja, Spain). PLOS ONE, 9(7), 49 e103634, doi:10.1371/journal.pone.0103634. 50 51 Vaquero, M., Van der Made, J., Blain, H. A., Ibánez, N., López-García, J. M., Rivals, F., et al. (2017). Fauna, 52 53 environment and human presence during MIS5 in theNorth of Spain: The new site of Valdavara 3. 54 Comptes Rendus Palevol, doi:10.1016/j.crpv.2018.03.004. 55 56 Varanda, A. (2018). O Acheulense de Grandes Lascas no Centro Sul de Portugal. Universidade de Lisboa, 57 58 59 Vázquez Varela, J. M. (1973). Contribución al estudio del paleolítico gallego. La comarca de río Arenteiro. 60 Cuadernos de Estudios Gallegos, XXVIII, 13-22. 61 62 22 63 64 65 Vialet, A., Modesto-Mata, M., Martinon-Torres, M., Martinez de Pinillos, M., & Bermudez de Castro, J. M. (2018). A reassessment of the Montmaurin-La Niche mandible (Haute Garonne, France) in the context of European Pleistocene human evolution. PLOS ONE, 13(1), e0189714, doi:10.1371/journal.pone.0189714. 1 2 3 Viana, A. (1930). Estações Paleolíticas do Alto Minho. Portucale, III(15), 5-51. 4 5 Villa, P., Soriano, S., Grün, R., Marra, F., Nomade, S., Pereira, A., et al. (2016). The Acheulian and Early 6 Middle Paleolithic in Latium (Italy): Stability and Innovation. PLOS ONE, 11(8), e0160516, 7 8 doi:10.1371/journal.pone.0160516. 9 10 Viveen, W., Braucher, R., Bourlès, D., Schoorl, J. M., Veldkamp, A., van Balen, R. T., et al. (2012). A 0.65Ma 11 chronology and incision rate assessment of the NW Iberian Miño River terraces based on 10Be and 12 13 luminescence dating. Global and Planetary Change, 94-95, 82-100, 14 doi:10.1016/j.gloplacha.2012.07.001. 15 16 Zbyszewski, G. (1943). La classification du paléolithique ancien et la chronologie du quaternaire de Portugal 17 en 1942. Boletin da Sociedade Geológica de Portugal, II(II-III), 3-112. 18 19 20 Zbyszewski, G. (1946). Etude géologique de la région d'Alpiarça. Comunicações dos servicios geologicos 21 de Portugal, XXVII, 125. 22 23

24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 23 63 64 65 FIGURE CAPTIONS Fig. 1 Chronological framework for main Middle Pleistocene sites in study area. MIS boundaries are from Lisiecki and Raymo (2005). 1 2 3 Fig. 2 Geographical (A) and geological (B) contexts for the sites cited in the text: 1: Louselas, 2: Valdavara 4 5 3, 3: As Lamas, 4: San Cibrán de Lás, 5: A Piteira, 6: A Regata, 7: Allariz´ surroundings, 8: Arbo, 9: Carvalhas, 6 10: Porto Maior, 11: Fillaboa, 12: Gándaras de Budiño, 13: Chan do Cereixo, 14: Marinho, 15: Cerro, 16: 7 8 Praia da Aguda, 17: Melhada, 18: Casal de Santa Maria, 19: Quinta do Cónego, 20: Casal do Azemel, 21: 9 10 Gruta da Aroeira/Galeria Pesada, 22: Ribeira da Ponte da Pedra, 23: Fonte da Moita, 24: Vale do Forno and 11 25: Monte Famaco. 12 13 14 Fig. 3 Examples of different LCTs and giant core of the Coastline area: a-c: Marinho site, d-f: Praia da Aguda 15 16 site and g: Cerro site 17 18 Fig. 4 Handaxes and cleavers on flake of different sites in the Miño River basin: a-d: Arbo site, e-h: Gándaras 19 20 de Budiño site 21 22 23 Fig. 5 Some examples of handaxes and cleaver on flakes of Porto Maior (a-d) and As Carvalhas (e-i) sites 24 25 26 Fig. 6 Handaxes and cleavers on flake of the Lis River basin: a-b: Pousias-Quinta do Cónego site, c-f: Casal 27 do Azemel and g-h: Casal de Santa Maria 1 site 28 29 30 Table 1 Summary of the data available for the most relevant acheulean sites mentioned in the text 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 24 63 64 65 Table Click here to access/download;Table;Tables.docx

Faunal Excavation Lithic % lithic Site Country Chronology Geological context LCTs % LCTs remains area (m2) remains m2 Louselas Spain Middle Pleistocene Holocene reworked soil No - 502 139 27.7 - As Lamas Spain Middle Pleistocene Fluvial terrace/Colluvium No 12 326 36 11.0 27.2 San Cibrán de Lás Spain Middle Pleistocene Holocene reworked soil No - 475 65 13.7 - A Piteira Spain Middle-Upper Pleistocene Alluvial fan No - 1320 - - - A Regata Spain Middle Pleistocene Fluvial environments No - 246 71 28.9 - Allariz´ surroundings Spain Middle Pleistocene Fluvial environments No 20 ~1500 - - - Arbo Spain >MIS 5 Alluvial fan No 30 3142 214 6.8 104.7 Carvalhas Portugal Middle Pleistocene Fluvial terrace No 42 439 38 8.7 10.5 Porto Maior PM4 Spain MIS 7 Fluvial terrace No 11.8 159 111 69.8 13.5 Gándaras de Budiño Spain Middle Pleistocene Fluvial terrace No 251 3759 218 5.8 15.0 Chan do Cereixo Spain Middle Pleistocene Holocene reworked soil No - 405 67 16.5 - Marinho Portugal Middle Pleistocene Colluvium No 104 2259 31 1.4 21.7 O Cerro Portugal Middle Pleistocene Colluvium No 22 199 57 28.6 9.0 Praia da Aguda Portugal Middle Pleistocene Marine terrace No - 154 4 2.6 - Casal de Santa María 1 Portugal Middle Pleistocene Colluvium No 52 336 28 8.3 6.5 Pousias-Quinta do Cónego Portugal Middle Pleistocene Fluvial terrace No 7 886 230 26.0 - Casal do Azemel Portugal Middle Pleistocene Colluvium No 151 3957 683 17.3 26.2 Gruta da Aroeira/Galeria Pesada Portugal MIS 11-8 Karstic cave infill Yes 28 1376 73 5.3 49.1 Ribeira da Ponte da Pedra Portugal MIS 8-7 Fluvial terrace No 21 1014 2 0.2 48.3 Fonte da Moita Portugal MIS 8-7 Fluvial terrace No 50 2852 1 0.0 57.0 Vale do Forno 1 Portugal MIS 9-7 Fluvial terrace No 25.5 246 32 13.0 9.6 Vale do Forno 3 Portugal MIS 6 Colluvium No 13 338 27 8.0 - Vale do Forno 8 Portugal MIS 9-7 Fluvial terrace No 25.5 3000 - - 117.6

Table 1

Figure 1 Click here to access/download;Figure;Fig-01.jpg Figure 2 Click here to access/download;Figure;Fig-02.jpg Figure 3 Click here to access/download;Figure;Fig-03.jpg Figure 4 Click here to access/download;Figure;Fig-04.jpg Figure 5 Click here to access/download;Figure;Fig-05.jpg Figure 6 Click here to access/download;Figure;Fig-06.jpg