UDK 903.2(36)"633/634":551-583:314-I8 Documenta Praehistorica XXXVI (2009) Climate change and population dynamics during the Late Mesolithic and the Neolithic transition in Iberia Javier Fernandez Lopez de Pablo1, Magdalena Gomez Puche2 1 Department of Anthropology, University of California, Santa Barbara, US javierfernadez@anth.ucsb.edu 2 Departamento de Prehistoria y Arqueolog^a, Universidad de Valencia, ES gomezpuc@uv.es ABSTRACT - This paper explores how Early Holocene climate changes in the Western Mediterranean would have affected Late Mesolithic settlement distribution and subsistence strategies in Iberian Peninsula, thereby giving rise to various adaptive scenarios. The current radiocarbon data set concerning the Neolithisation process has revealed the rapidity of the spread of farming in Iberia. Considering both the implications of the last hunter-gatherers' adaptation strategies and the population dynamics of agro-pastoral communities, we address the migration patterns underlying the Mesolithic-Neolithic transition. In conclusion, we propose that the initial colonization process was the result of two successive and spatially heterogeneous migrations: Maritime Pioneer Colonization and targeted migration to places favorable to the new economic system. IZVLEČEK - V članku raziskujemo kako bi zgodnje holocenske klimatske spremembe v zahodnem Sredozemlju vplivale na pozno mezolitsko razporeditev naselbin in na prehranjevalne strategije na Iberskem polotoku in s tem povzročile nastanek različnih prilagoditvenih scenarijev. Trenutna serija radiokarbonskih datumov, povezanih s procesom neolitizacije je razkrila hitrost širjenja poljedelstva na Iberskem polotoku. Ob upoštevanju obojega - vpletenosti prilagoditvenih strategij zadnjih lovcev in nabiralcev ter populacijske dinamike poljedelsko-pastirskih skupnosti, opazujemo selitvene vzorce, ki so podlaga mezolitsko - neolitskega prehoda. V zaključku predlagamo, da je bil začetni proces kolonizacije rezultat dveh zaporednih in prostorsko heterogenih migracij: pomorske pionirske kolonizacije in ciljne migracije na prostore naklonjene novemu ekonomskemu sistemu. KEY WORDS - climate change; 8200 calBP event; Mesolithic-Neolithic transition; migration Introduction As elsewhere in Europe, the origin of the Neolithic in Iberia is related to the introduction of technical and economic innovations in food production by means of different migration processes and subsequent cultural interaction with local populations of hunter-gatherers (Zilhao 2001; Bernabeu 2002). Consequently, one can expect a regionally diverse and complex process modeled by both environmental and social factors. In the 90's, discussions on the Neolithic transition in Iberia were framed within the migrationist and in- digenist debate (e.g. Bernabeu 1996; Vicent 1997) and the refutation of polygenist models (Zilhao 1993; Bernabeu et al. 1999). However, during the last 10 years, the chronology of the first ceramic contexts and domesticates have enjoyed renewed interest as a result of the incorporation of new data from the Iberian interior and the Northern Façade (see below). In the context of the whole of Iberia, both regions, as counterparts of other traditional research areas such as eastern Spain and Portugal, focus most of the current debate on the timing and mechanisms behind the farming expansion (e.g. Al- DOI: 10.4312/dp.36.4 67 day 2005; Bernabeu 2007; Cruz and Vicent 2007; Carvalho 2003; Juan-Cabanilles and Marti 2002; Marti 2008; Rojo et al. 2006 among others). In this context, the relationship between environmental and cultural dynamics regarding the Neolithic transition has been placed on a secondary level of analysis and interpretation. With the exception of Zilhao, who has openly argued how these factors explain the absence of a Mesolithic population in the central hinterlands of Portugal and Spain (Zilhao 2000; 2001), very little attention has been paid to this issue until recently. Nevertheless, new paleo-environmental evidence and archaeological data on the Late Mesolithic record seem to indicate that some transformations in the settlement and subsistence patterns erupted suddenly: 1. A geographically circumscribed distribution of Mesolithic settlement. There is still a lack of archaeological information regarding the Late Mesolithic from several areas, such as Spanish Meseta and Catalonia, and research bias alone cannot explain this. 2. Variable reliance on aquatic resources shown among the last groups of hunter-gatherers. The available information on paleodiets and isotopic analyses of Mesolithic populations from Portugal, Canta-brian Façade and the central Mediterranean coast of Spain, have provided new data of relevance to this issue, reflecting regional disparities. In this sense, the degree of dependence of the Mediterranean samples is remarkably inferior to those recorded in Portuguese and Cantabrian shell middens. From an evolutionary perspective, if the first point can be linked to the dynamics of environmental change and to the adjustments that Mesolithic population made, the second is a direct outcome of these adaptations. This paper will explore how Early and Middle Holo-cene environmental changes could have affected the geographical distribution, organization and subsistence of Late Mesolithic settlements. This period witnessed the appearance of the Neolithic in Iberia, providing various opportunities for ongoing farming dispersal processes in the western Mediterranean. A critical review of the last foragers evidence in Iberia is essential to establish a coherent frame of hypotheses about the role of indigenous populations in the economic and demographic changes that occurred during the Neolithic transition. The paper is in two parts. First, we present a general overview of Early-Middle Holocene environmental dynamics and Late Mesolithic settlement distribution in Iberian Peninsula. Second, we discuss the effects of the 8200 calBP event on regional settlement organization and sedimentation dynamics. Finally, we suggest how these dynamics would have affected subsistence patterns on the basis of paleo-dietary studies. In the second part, we present empirical evidence from the earliest Neolithic sites from the latest regional studies in the Iberian bibliography. We then revisit different models of the transition to farming in the Iberian Peninsula, prior to sketching new interpretations that emphasize population dynamics and environmental changes alongside the Mesoli-thic-Neolithic transition. Also, since the Neolithisa-tion process entailed complex demographic transformations, we discuss paleogenetics and population replacement regarding the Iberian data. Finally, we propose some directions for future research. The body of Mesolithic and Neolithic radiocarbon dates for the Iberian Peninsula has been particularly increased and enhanced in recent times due to the systematic application of sample selection protocols Juan-Cabanilles and Marti 2002; Rojo et al. 2006; Bernabeu 2006; Carvalho in press). The current compilation - in Tables 1 and 3 - has been built on the basis of a series of radiocarbon dates from single, short-lived samples: cultivated plants (mainly cereals), non-domesticated short-lived fruits (acorns), domestic fauna and human bones, following Venice's 1998 conference recommendations (Ammerman and Biagi 2003). Individual AMS 14C dating of key specimens overrides the risk of dealing with disturbed contexts and with intrusions from overlying levels (Bernabeu et al. 1999; Zilhao 2001) and eliminates the possibility of the 'old wood' effect in the case of charcoal (Zilhao 2001; Zapata et al. 2004. 285). In addition, we have excluded from this analysis the radiocarbon determinations on shell samples - although the correction can be determined, their reservoir effect values, locally and diachroni-cally, are subject to considerable variation (e.g. Soa-res and Dias 2006) and have not been established on the basis of short-lived samples from Neolithic contexts. Consequently, an unpredictable degree of uncertainty affects radiocarbon determinations on shell samples, and do not allow a comparisons with other Iberian contexts. On the other hand, we have included the radiocarbon dates of human remains from the Muge sites (Cabeço da Arruda, Cabeço da Amoreira and Moita do Sebastiao), calibrated considering several potentially changeable factors such as the percentage of marine resources consumed revealed by isotopic analyses, and the local estuarine reservoir effect (Martins et al. 2008), which is different from that established by Soares (1993)1 Early-Middle Holocene environmental changes and Late Mesolithic distribution In current debates on the Neolithic transition in Iberia, the distribution of the last hunter-gatherer populations during the Neolithisation process is one of the main issues. In the last decade, Iberian archaeology has witnessed an outstanding advance in knowledge of the Late Mesolithic, which has changed the traditional archaeological sequence (Fortea 1973). Today, the Mesolithic in the Iberian Peninsula (c. 10 800-7200 calBP) is comprised of two successive industrial complexes that led regionally different cultural traditions: the flake-rich assemblages complex that dominated during the Boreal period and the Geometric Mesolithic, also generally called the Late Mesolithic, the main cultural features of which are presented below: • Flake rich assemblages complex: The main distinguishing feature of this complex is a lithic industry based on flake technology with no - or very little -evidence of blade production. Lithic assemblages vary from flint to quartzite, but flake artifacts, especially notches and denticulates, and massive tools, comprise all of them. There are several denominations in the archaeological literature to define this kind of lithic industry (Mesolithic Macrolithic or Generic Me-solithic), including archaeological entities such as the 'Notches and Denticulates Mesolithic' (henceforth ND Mesolithic) (mainly in the Mediterranean region, the Ebro Valley and the Pyrenees) (Alday 2006a; Cava 2004) or Asturian (see Straus 2008 for a recent revision). This complex dates between 10 200 and 8400 calBP, except the Asturian in the western Cantabrian region, where this complex is not well distinguished in typological terms from the Late Mesolithic phase due to the scarcity of geometric microliths at many open air sites. • Late Mesolithic phase A: During this phase of the Iberian Mesolithic, there is a marked technological change with the re-introduction of blade debitage technology, the microburin technique and the configuration of trapezoidal microliths with abrupt retouch. This phase lasts from 8400 to 7900 calBP. • Late Mesolithic phase B: Considered an evolution from the previous phase, it is characterized by the presence of triangles among the geometric micro-liths. In this sense, the most outstanding phenomenon is the emergence of specific microlith types with a regionally discrete distribution: triangles with concave sides known as 'Cocina triangles' at Valencia and Aragon regions, 'Sonchamp points' in the western Pyrenees and 'Muge triangles' in Portugal very similar to their Cocina correlatives. This phase lasts from 7900 calBP up to the beginning of the Neolithic in the different regions of Mediterranean Iberia (7500-7200 calBP). • Late Mesolithic phase C: this is considered the terminal development of the Mesolithic industries that paralleled the expansion of the Early Neolithic (Fortea 1973; Juan-Cabanilles 1990). It is characterized by triangles and segments with bifacial retouch. However, its Mesolithic cultural affiliation is now subject to review, given the recent documentation of this set of geometric armatures from the beginnings of the Neolithic at some Cardial and Epicardial sites such as Chaves Cave (Cava 2000) with no underlying layers of Mesolithic occupation. In fact, the identification of Phase C, in typology, technology, stratigraphy and absolute chronology is ambiguous and not clearly isolated from earlier or later occupations (Juan Ca-banilles and Marti 2007-2008). As noted for other southern European regions (Biagi 2003; Binder 2000; Juan-Cabanilles and Marti 2002; Carvalho in press), considerations about the Late Mesolithic in terms of social geography and population dynamics should be grounded on a critical evaluation of chronology and the archaeological evidence, i.e. on the radiocarbon framework and the li-thic industries. According to this, the map in Figure 1 gives an accurate picture of Late Mesolithic distribution (phases A and B) in the Iberian Peninsula (Fig. 1). It is based on a complete compilation that encompasses different kinds of archaeological site (rock-shelters, open-air sites and lithic scatters), with radiocarbon dates or accurate typological information published until 2002 (Juan-Cabanilles and 1 Calibrations, as well as the corresponding graphs, were obtained using the Version 4.1. of the OxCal Program (Bronk-Ramsey 2009), based on the IntCal04 curve (Reimer et al. 2005). All radiocarbon dates mentioned are in years BP and BC after calibration, and based on extremes of the 2 sigma range. Fig. 1. Regional distribution of Late Mesolithic sites in Iberian Peninsula. 1. Llatas; 2. Gelat.; 3. Falguera; 4.Vacas; 5. Polvorosa; 6. Cocina; 7. Anna; 8. Tossal de la Roca; 9. Collado; 10. Casa de Lara; 11. Huesa Tacana; 12. Mas Nou; 13. Mas de Martv 14. Mas de Sanç; 15. Ballester; 16. Cavall; 17. Estany; 18. Santa Maira; 19. Encantada; 20. Regadiuet; 21. Peneta; 22. Ceja; 23. Mas Cremat; 24. Zorra; 25. Sitjar; 26. As-sud; 27. La Mangranera; 28. Nerja; 29. Nacimiento; 30. Valdecuevas; 31. Frailes; 32. Retamar; 33. Rio Palmones; 34. Roca das Gaviotas; 35. Castelejo; 36. MontesBaixo; 37. Fiais; 38. Medo Tojeiro; 39. Vidi-gal; 40. Samouqueira; 41. Vale do Pincel; 42. Forno Telhas; 43. Forno da Cal; 44. Buraca Grande; 45. Ara-pouco; 46. Rebolador; 47. P. Sao Bento; 48. Muge region (Cabeço da Amoreira, Cabeço da Arruda, Moita do Sebastiäo); 49. Abrigo Bocas; 50. Reiro; 51. Xestido III; 52. Espertm ; 53. Colomba; 54. Coberizas; 55. La Riera; 56. Sierra Plana; 57. Mazaculos; 58. Los Canes; 59. Covajorno; 60. El Aguila; 61. Pendueles; 62. Toralete; 63. La Garma A; 64. Truchiro; 65. CuMo Redondo; 66. Cofresnedo; 67. La Chora; 68. Tarreron; 69. La Fragua; 70. El Perro; 71. La Trecha; 72. Pareko Landa; 73. Urratxa; 74. Linatzeta; 75. Herriko Barra; 76. Marizulo; 77. Monticocha; 78. Pena Maranon; 79; Padre Areso; 80. Socuevas; 81. Angel; 82. Vidre; 83. Costalena; 84. Salada Grande; 85. Pontet; 86. Sol Pinera; 87. Serda; 88. Cabezo Cruz; 89. Mendandia; 90. Kanpanoste G.; 91. Atxoste; 92. Fuente Hoz; 93. Los Banos; 94. Botiqueria; 95. Forcas II; 96. Lagrimal; 97. Aizpea; 98. Margineda; 99. Bajondillo. Marti 2002). We have updated it by adding new data, published since then (see below). The body of archaeological data displays clear evidence of occupation in the Mediterranean Façade, the Ebro Valley, the Cantabrian Façade and central and south Portugal. Of course, this picture results from different regional research trajectories, which have produced quantitatively and qualitatively diverse information. However, the current distribution pattern does not differ substantially from that published seven years ago. The central Mediterranean area (the Valencia region) has a significant number of Late Mesolithic sites (for a recent revision, see Aura et al. in press). Recent work has discarded previous Late Mesolithic attributions, such as the site of Arenal de la Virgen (Fernandez et al. 2008), but added some new sites, such as Mas de Marti (Fernandez et al. 2005), Santa Maira (Aura et al. 2006), Mas Gelat (Miret el al. 2007), Mas de Sanç (Fernandez 2006), Cueva del Lagrimal (Gomez and Fernandez 2009) and Mas Cremat (Vicente et al. 2009). Moreover, a previously known site, Falguera, has recently been published in full (Garcia and Aura 2006). As a result of continuous and intensive archaeological research, the Ebro Valley provides an outstanding Late Mesolithic record (Alday 2002; Barandia- SITES SAMPLE CONTEXT MATERIAL SAMPLE REFERENCE 14C age (BP) CalBP 20 range CalBC 20 range Ref. PORTUGAL Cabeço da Arruda Skeleton N H Homo TO-356 6360 ± 80 7170-6680 5220-4730 1 Cabeço da Arruda Skeleton D H Homo TO-355 6780 ± 80 7690-7370 5740-5420 1 Cabeço da Arruda Skeleton 42 H Homo TO-359a 6960± 60 7790-7440 5840-5490 1 Cabeço da Arruda Skeleton A H Homo TO-354 6970 ± 60 7860-7560 5910-5610 1 Cabeço da Arruda Skeleton III H Homo TO-360 6990 ± 110 7930-7450 5980-5500 1 Cabeço da Arruda CA-00-01 H Homo TO-10217 6620 ± 60 7520-7230 5570-5280 1 Cabeço da Arruda CA-00-02 H Homo TO-10216 7040 ± 60 7870-7570 5920-5620 1 Fiais - B Bone ICEN-141 6180± 110 7233-6913 (1s) 5400-4800 2 Fiais S.XIX, A10, z.244 B Mammal TO-706 6260 ± 80 7126-7026 (1s) 5470-5000 2 Fiais - B Bone ICEN-110 6870± 220 8160-7330 6250-5350 2 Moita do Sebastiao Skeleton CT H Homo TO-135 6810 ± 70 7570-7250 5620-5300 1 Moita do Sebastiao Skeleton 41 H Homo TO-134 7160 ± 80 7940-7590 5990-5640 1 Moita do Sebastiao Skeleton 24 H Homo TO-132 7180 ± 70 7950-7620 6000-5670 1 Moita do Sebastiao Skeleton 29 H Homo TO-133 7200 ± 70 7970-7640 6020-5690 1 Moita do Sebastiao Skeleton 22 H Homo TO-131 7240 ± 70 7980-7640 6030-5690 1 Samouqueira I c.2 H Homo TO-130 6370 ± 70 6800-6633 (1s) 5480-5210 2 Vidigal c.2 (shell midden) B Bones GX-145557 6030 ±180 7439-6305 5490-4356 3 Vidigal c.3 (paving) B Bones Ly-4695 6640 ± 90 7738-7274 5789-5325 3 Cabeço da Amoreira Skeleton ? H Homo TO-11819R 7300 ± 80 8050-7660 6100-5710 1 Cabeço da Amoreira Skeleton 7 H Homo Beta-127450 6850 ± 40 7610-7380 5660-5430 1 Cabeço da Amoreira CAM-00-01 H Homo TO-10218 6630 ± 60 746°-717° 5510-5220 1 Cabeço da Amoreira CAM-01-01 (139) H Homo TO-10225 6550 ± 70 7980-7640 5630-5370 1 Vale de Boi c.2 H Homo TO-12197 7500 ± 90 8551-8020 6602-6071 4 CANTABRIAN STRIP AND PYRENEES Linatzeta - H Homo KIA-33193 7315 ± 35 8300-8002 6351-6053 5 Los Canes Skeleton 6-III H Homo AA-6071 6930 ± 95 8015-7518 6066-5569 6 Los Canes 6-II feet H Homo AA-5295 6860 ± 65 7932-7566 5983-5617 6 Los Canes Skeleton 6-II H Homo AA-5296 6770 ± 65 7826-7460 5877-5511 6 Los Canes Skeleton 6-II H Homo AA-11744 7025 ± 80 8052-7615 6103-5666 6 Los Canes Skeleton 6-I H Homo AA-5294 6265 ± 75 7424-6904 5475-4955 6 Colomba Shell midden H Homo TO-10223 7090 ± 60 - 5910-5534 7 Cubio Redondo Shell midden B Cervus Beta-106050 6630 ± 50 7622-7421 5673-5472 8 Cofresnedo Cof.5 B Roe-deer GrA-20146 6865 ± 45 7847-7581 5898-5632 7 Urratxa Level fertile B Bone Ua-11435 6995 ± 80 8025-7607 6076-5658 7 Urratxa Level fertile B Bone Ua-11434 6940 ± 75 7969-7585 6020-5636 7 MEDITERRANEAN REGION AND EBRO VALLEY El Collado Burial 12 H Homo UBAR-281 7640 ±120 8989-8050 7040-6101 9 El Collado Burial 12 H Homo UBAR-280 7570 ±180 9033-7858 7084-5909 9 Mas Cremat Level_III F Sorbus Beta-232342 6780 ± 50 7787-7507 5838-5558 10 Cingle Mas Nou ent - H Homo Beta-170715 6920 ± 40 7929-7623 5980-5674 11 Cingle Mas Nou 3 - H Homo Beta-170714 6910 ± 40 7924-7620 5975-5671 11 Falguera - S Olea sp. AA-2295 7410 ± 70 8404-8013 6455-6064 12 Lagrimal Level_IV B Ibex Beta-249933 6990 ± 50 7960-7676 6011-5727 13 Botiquerfa Level_4 B Bone GrA-13267 6830 ± 50 7835-7568 5734-5663 14 Botiqueria Level_2 B Bone GrA.13265 7600 ± 50 8554-8205 6605-6256 14 Aizpea Level_B B Bones GrA-779 6600 ± 50 76133-7335 5664-5386 15 Tab. 1. Late Mesolithic radiocarbon dates on short lived samples. H = human bone; B = bone; F = fruit; S = seed. References. 1: Martins et al 2008; 2: Lubell et al. 2007; 3: Carvalho in press; 4: Carvalho et al. 2008; 5: Tapia et al. 2008; 6: Arias 2005/2006; 7: Fano 2004; 8: Ruiz and Smith 2001; 9: Guixé et al. 2006; 10: Vicente et al. 2009; 11: Olària and Gusi 2005; 12: Barton et al. 1990; 13: Inedit; 14: Barandiaran and Cava 2000; 15: Barandiaran and Cava 2001. rân and Cava 2000; Cava 2004; Utrilla et al. 1998). Sites extend from the Lower Aragon region, with a clear spatial continuity into north Valencia, through the High Ebro Valley and the central and western Pre-Pyrenees. Recently, some new sites has been published in full, such as Mendandia (Alday 2006b) or Los Banos (Utrilla and Rodanés 2004), and some new Late Mesolithic sites with radiocarbon dates have been reported, such as Cova del Vidre (Bosch 2008) and Cabezo de la Cruz (Picazo and Rodanés 2008). The latter site fills the gap in the Late Meso-lithic record in the south central area of the Ebro Valley. In Andalusia, in southern Iberia, the Late Mesolithic remains understudied. The references traditionally cited are some interior rock shelter sites, such as Na-cimiento and Valdecuevas, or possible lithic scatters like Los Frailes (Juan-Cabanilles and Marti 2002). In recent years, a few open-air Late Mesolithic sites have been published in full for the Algeciras Bay area, such as Embarcadero del Rio Palmones (Ramos and Castaneda 2005) and El Retamar, on the Atlantic coast of Cadiz (Ramos et al. 2002). El Retamar is considered Neolithic by its excavator (Ramos et al. 2005), although the lithic assemblage - overwhelmingly dominated by trapezoidal microliths and microburins - and the radiocarbon dates suggest the existence of a preceramic occupation phase2. At the Malaga coast, the recent revision of two long cave sequences - Nerja and Bajondillo caves- has yielded Late Mesolithic evidence (Aura et al. 2005; Cortés 2007, respectively). In Portugal, the main clusters of Late Mesolithic sites are located around the Lower Tagus, Sado and Mira estuaries and the Alemtejo coastline (for regional syntheses, see Bicho 1994; Zilhäo 2000; Carvalho 2002; 2003). A secondary cluster is documented in the Rio Maior at the Estremadura region (Forno da Telha and Abrigo das Bocas), which is interpreted in terms of logistic dependence on the Muge Mesolithic sites (Carvalho 2003). Recently, new research projects have added some New Late Mesolithic sites in the Algarve (Carvalho et al. 2005; 2008; Stiner et al. 2003). To the north, in the Alto Douro, only the Prazo site has produced possible evidence of Late Mesolithic occupation; however, in the light of the provenance and nature of the radiocarbon dates (charcoal samples), its Mesolithic attribution is not unanimously accepted (for a detailed discussion, see Carvalho 2003; Zilhäo 2003; Monteiro-Rodrigues 2003). Along the north Iberian coast, some dispersed Late Mesolithic sites have been reported in Galicia (Vâz-quez 2004); however, the main archaeological evidence comes from an area between the eastern half of Asturias and the Basque Country (Fano 2004; Straus 2004; 2008). Previous approaches to this region traditionally outlined the importance of coastal adaptations in the settlement distribution with the abandonment of the mountain inland after the Azi-lian. However, archaeological research in recent years has changed this perspective. Late Mesolithic occupations are also documented for inland locations, such as Los Canes Cave (Arias 2005), and even in high mountain areas such as Espertin (Fuertes and Neira 2006). Also, some Late Mesolithic sites (most of them shell middens) dated to 7700-6600 BP have been reported recently, such as Linatzeta (Tapia et al. 2008), Cubio Redondo (Ruiz and Smith 2001), Colomba (Arias 2005/2006), Truchiro, Toralete, Cobe-rizas, Covajorno, El Aguila, Pendueles, and Sierra Plana (Fano 2004). Complete publications of new Late Mesolithic sites include Cofresnedo (Ruiz and Smith 2003). In contrast to the regions mentioned above, there is a lack of Late Mesolithic data in two large Iberian regions: the central region, the Tertiary plateau known as the Spanish Meseta, and Catalonia in northeast Spain. The interpretation of this 'archaeological silence' requires detailed discussion, using both paleo-environmental evidence and the representativeness of the archaeological research. Some scholars have suggested that the human population of the Iberian Peninsula interior during the Early and Middle Holo-cene was rather low until the beginning of the Neolithic (Zilhäo 2000.144; Straus 2008). Early Holocene reforestation would have produced a reduction in herbivorous biomass, making the area less attractive relative to coastal and estuarine areas. Consequently, scattered settlement or very low population densities lead to a poor archaeological record. At present, this interpretation is still supported by current paleo-en-vironmental and archaeological evidence. Effectively, North Meseta pollen records indicate a forest re-colonization from the beginning of the Ho- 2 Radiocarbon dates from this site also display a disparity with different statistical result. For instance, a single feature -hearth 18-provides two radiocarbon dates elaborated on shell samples with a difference of 500 years (Sac 1525: 6900±70 BP and Beta-90122: 6400±85 BP); the oldest one is statistically similar to another one coming from a different sector of the same site (Sac-7020±100 BP), which fit within the chronological limits of the Late Mesolithic. locene, and the progressive replacement of pine formations by oak forests. During the Boreal period, forest expansion reached its maximum extent, as reflected in both the Sanabria reservoir and the San-guijuelas Lagoon pollen diagrams, where tree pollen dominated by Quercus comprises 70% and 90% respectively (McKeever 1984; Munoz et al. 2001). During this period of forest expansion, the number of archaeological sites declines drastically. Final Paleolithic sites with Magdalenian industries are well known in the North Meseta (Diez and Delibes 2006). In contrast, the number of Preboreal sites decreases significantly, and there is no (stratigraphic, chronological or industrial) reliable evidence of Boreal sites or Late Mesolithic sites from North Meseta (Alday 2005; Corchon 2006). In this sense, some presumed Mesolithic attributions published recently need to be revisited. For example, Jiménez-Guijarro (2005 and 2008) has pointed out the presence of Late Mesolithic occupations in the Madrid region at the Ventana Cave and at the open air site of Verona II. However, the published archaeological evidence lacks radiocarbon dates, stratigraphy or diagnostic artifacts to confirm such an attribution3. In addition, other archaeological deposits considered by this author as Late Mesolithic, such as Verdelpino or the Nispero caves (Jiménez-Guijarro 1999) are not Late Mesolithic, but Epipaleo-lithic, dating to the Early Holocene. There is still a notable lack of Late Mesolithic, especially considering that systematic extensive survey programs have been developed in the North Meseta (Balbin et al. 1997; Rojo and Kunst 1999). Furthermore, several Holocene archaeological sequences have been excavated in recent years, such as El Mirador Cave (Vergès et al. 2008), El Portalon Cave (Ortega et al. 2008), and the Carlos Alvarez rock shelter (Rojo et al. 2008), but no Late Mesolithic levels have been reported. Catalonia, in north-east Spain, is another second region where no Late Mesolithic sites have been documented. This lack of data does not seem to be the result of bias in archaeological research. Early Holo-cene occupations with Epipaleolithic and Sauveter-rian industries on the central coast of Catalonia and in the Interior Valleys of Lerida (sites like Parco, Fila- dor, and Balma de Gai) (Garcia-Argüelles 2006) are well documented. In addition, many Boreal ND Meso-lithic sites are documented for the same areas (Par-co and Filador) (Garcia-Argüelles 2006), the south east Pyrenean foothills (Martinez-Moreno 2006) and in central and southern Catalonia (Vaquero 2006). The last Mesolithic occupations documented in Catalonia are from the ND Mesolithic complex - level A of the rock-shelter at Cativera - dating to 9000-8600 calBC (Allué et al. 2000). A gap of more than 1200 years separates the most recent evidence of foragers from the earliest Cardial Neolithic context. The lack of Late Mesolithic sites in Catalonia might be related to a broader process of settlement reorganization documented in the Mediterranean at the end of the Boreal period (Fernandez and Jochim in press). In this sense, it should be noted that the last ND Mesolithic occupations documented in Catalonia - the Cativera and Moli del Salt sites - are recorded in the upper section of their stratigraphic sequences, while in the neighboring central Ebro Valley, the ND Mesolithic inaugurates the occupation of many rock shelters which display a clear occupational continuity thereafter, during the Late Mesolithic (Montes et al. 2006). The settlement reorganization implies a general increase in the number of Late Mesolithic sites, as well as changes in the settlement pattern, with a significant increase in open-air sites around inland and coastal lagoons. Necropolises, such as El Collado (Aparicio 1990; Garcia Guixé et al. 2006) or Mas Nou (Olària et al. 2005), completely absent in the ND Mesolithic archaeological record, are another innovation of this period. In Portugal, the Late Mesolithic witnessed the emergence of year-round residential campsites located in highly productive estuarine areas. The most visible manifestations from the Atlantic period are shell middens with a complex interior spatial organization that included a number of post-holes and pit structures, such as roasting and storage pits, hearths, etc, and necropolises formed by numerous inhumations (i.e. Moita do Sebastiao) (Roche 1972). Regarding the Boreal period, the settlement pattern was reorganized, and sites are clustered only in estuarine areas, leaving inland territories, like the Estremadura plateau case, uninhabited (Araujo 2003). 3 At the Ventana Cave, the presence of trapezes with bifacial retouch and the length and width of the trapezes with abrupt retouch is quite similar to the microlith assemblages founded in Megalithic tombs during the Middle Neolithic (Alegre 2005; Fernandez et al. 2009). Environmental dynamics and Late Mesolithic settlement organization: the 8200 calBP event The 8200 calBP event in southern Europe is currently being considered as the main environmental factor triggering the dispersal of agriculture (Weninger et al. 2006; Budja 2007), land use changes (Gonzâ-lez-Sampériz et al. 2008; Fernândez and Jochim in press), or the cause of the chronological gaps observed in the radiocarbon record between the Late Me-solithic and the Early Neolithic periods (Berger and Guilaine 2009). In contrast to the eastern Mediterranean, the 8200 calBP event in Iberia fell between Late Mesolithic phases A and B. Recently, scholars have used its effects to explain changes in land use patterns that potentially affected both the regional distribution and settlement organization of the last forager populations in Portugal (Zilhäo 2003; Carvalho in press), the Ebro Valley (Utrilla 2005; Gonzâlez-Sampériz et al. 2008) and the Mediterranean region (Fernân-dez and Jochim in press). A different issue turns on the potential relationship between this climatic event and sedimentation dynamics, as might be suggested by the increasing documentation of sites having a stratigraphic hiatus between the Late and the Early Neolithic occupations (Berger and Guilaine 2009). In Portugal, the relationship between the 8200 calBP event and the decline in marine productivity has been pointed out to explain the habitat concentration on the large estuarine areas such as the Lower Tagus and the Sado (Zilhäo 2003; Carvalho in press). According to this hypothesis, this climatic episode allowed a freshwater current to reach the Portuguese coast, reducing coastal upwelling activity. The main indicator to support this relies on the diachronic variation in the reservoir effect along the Portuguese coast during the Holocene being considered as a coastal upwelling proxy (Soares and Dias 2006). However, the chronological correlation between coastal upwelling activity and changes in the settlement pattern is not completely clear due to the current chronological scale for determining the reservoir effect on the basis of radiocarbon dates of shell and charcoal samples. What seems clear from the archaeological evidence is the crucial effect of this climatic crisis on the abrupt change in coastal occupation and subsistence pat- terns documented from the Boreal to the Late Meso-lithic periods (Araujo 2003; Bicho 2006; Carvalho in press). For the former period, a higher number of small or medium size shell middens have been documented at different coastal locations in the Algarve, Alemtejo and Estremadura regions, while the Late Mesolithic witnessed habitat concentration around the main estuaries, with a significant increase in shell midden size. A lower degree of residential mobility, with mostly round year occupation, is reflected by the documentation of necropolises on the shell middens, and a higher dependence on aquatic resources, estimated at around the 40%-50% of the diet on the basis of the stable isotopic evidence (Lubell et al. 1994). In this sense, a strong chronological correlation between Lower Tagus estuarine adaptations and the beginning of the 8200 calBP event can be established if isotopic corrected human bone samples are considered on the basis of the local determination of the reservoir effect at the Muge region (Martins et al. 2008.83)4. In addition, around 8150 calBP, the development of the Lower Muge shell middens was enabled with an abrupt change from fluvial to estua-rine environments, which led the to formation of productive shell beds (van der Shrieck et al. 2008). In the Mediterranean climatic region of Iberian Peninsula, a significant number of paleo-environmental studies - including both marine and lake cores -have recorded short-term stepped variations in temperature, water balance input and vegetation during the 8200 calBP (Fig. 2). Isotopic analysis from three marine cores located in the Alboran sea (Cacho et al. 1999) and close to the Menorca island (Frigola et al. 2007; Jiménez-Espejo et al. 2007) have documented a downturn of 2° C in sea surface temperature. According to Frigola et al. (2007.13), this change was associated with a persistent positive index in the North Atlantic Oscillation that produced colder and drier conditions in the Iberian Peninsula (Fig. 2). Multi-proxy studies from several continental lakes show a chronological correlation with lower lake levels in La Estanya Lake in the pre-Pyrenean mountains (Morellôn et al. 2007.12-13) and in Salinas Playa Lake, more than 600km further south in southeast Spain (Giralt and Julià 2003.321). Vegetation changes have been documented in several natural and archaeological deposits in eastern Spain, indicating an interruption in the temperate 4 The old wood effect problems in radiocarbon dating on Muge shell middens have been broadly discussed in recent approaches (Zilhäo 2001; Jackes and Meicklejohn 2006; Roksandic 2006; van der Shriek et al. 2008). Fig. 2. Early to Middle Holocene environmental dynamics from different paleo-climatic proxies in the Mediterranean region of Iberia where the 8200 cal yr BP event is detected (source: Gonzalez-Sampériz et al. 2009.Fig.4 adapted). Juniperus and Mesophhytes percentages at Portalet lake sequence (Gonzalez-Sampériz et al. 2006); Sulphur (S) content (count per seconds) (Morellon et al. 2008); Minorca Drift record (Core MD99-2343, Mediterranean Sea) Potassium (K) -percentage (%) and Silicon/Aluminium (Si/Al) ratio (Frigola et al. 2007) and Greenland GISP2 ice core oxygen isotope curve (Grootes and Stui-ver 1997). A grey band marks the 8.2 ka event as referenced at the GISP2 ice core (Gro-otes and Stuiver 1997). and moister conditions that had resulted in forest expansion since the beginning of the Holocene (Lopez Sâez et al. 2007). During the 8200 calBP event, El Portalet Lake underwent an abrupt decrease in Juniperus, Betula, Corylus and deciduous Quercus taxa (Gonzâlez-Sampériz et al. 2006.49). Further south, the San Rafael sequence at Almeria also records an abrupt change from mesic to xeric vegetation (Lopez Sâez et al. 2007). In the Central Ebro Desert, multi-proxy lake records of Guallar and Hoya del Castillo lagoons show a stepped increase of microfossil charcoals around 8.3 ka BP caused by natural fires as a result of a general decline in annual rainfall. The increase in microfossil charcoal is also associated with vegetation changes, mainly the development of fire-resistant evergreen oaks such as hermes oak and, especially, cork oak, which replaced the pine forests (Davis and Stevenson 2007.1706). Finally, although pollen reconstructions from Late Mesolithic archaeological deposits in caves or rock shelters are deficient to document the 8200 calBP event, several sites such as the El Nacimiento or El Bajondillo caves saw an increase in xeric species (Lopez Sâez et al. 2007). In Northern Iberia, the paleo-environmental data required to detect the short-term vegetation changes produced by such climatic crisis are less informative. In the north west sector (Galicia), an episode of oak forest retreat has been recently correlated with the 8200 calBP event, changing the traditional interpretation by which the regression of forest biomass was caused by Epipaleolithic populations (Martinez-Corti-zas et al. 2009.82). In contrast, in the central sector of the northern Iberian Peninsula (the Western Pyrenees, East Cantabrian Mountains and the Upper Ebro Valley), no relevant vegetation changes have been documented for the 8th millenium, which displays a continuous expansion of deciduous forests and a progressive decline in pine formations (Iriarte 2009). Unfortunately, the archaeological information is biased, given the poor preservation of palynomorphs at relevant Late Mesolithic sites such as Mendandia, Kanpa-noste and Kanpanoste Goikoa (Iriarte 2009.70). The ecological impact and its consequences on human land-use patterns in the Mediterranean region, especially in the Central Ebro Valley, are the subject of lively debate. The climate in this particular area is Mediterranean, with a continental influence, resulting in semi-arid conditions manifested by hot summers, cold winters, low annual precipitation and high evapo-transpiration index. Some scholars, like P. Utrilla and P. Gonzalez-Sampériz, have correlated the lack of Mesolithic sites dated c. 7300-6800 BP at the Bajo Aragon with this aridity crisis. This chronological gap is related to the documentation of sterile levels in the main Mesolithic rock shelters of the area, such as Botiqueria, Los Banos, Angel, and Pontet. These authors argue that these arid conditions compelled people to move to other, wetter areas such as the High Ebro Valley, the eastern, central Pyrenees, or the Maestrazgo (Utrilla 2005; Gonzâlez-Sampé-riz et al. 2008). Consequently, Lower Aragon witnessed no human occupation or very low occupation intensity until the beginning of the Neolithic. Nevertheless, other authors have recently argued for a shorter chronological influence, considering an individual evaluation of the stratigraphy and the radiocarbon dates (Fernândez and Jochim in press). Furthermore, similar situations of site abandonment might potentially be identified in other Iberian areas under Mediterranean-continental influence, such as South Valencia or the central Pre-Pyrenees foothills. Effectively, sites such as Forcas II and Falguera also document sterile layers formed during part of the 8200 calBP climate crisis bounded by Mesolithic occupations (Fernândez and Jochim in press). However, the main discussion point relies on the human responses to such abandonment episodes. Utrilla and Gonzalez-Sampériz interpret the sterile levels in terms of population dynamics, arguing for a depopulation of Lower Aragon and subsequent migration to other, wetter areas (Utrilla 2005; Gonzâlez-Sampériz et al. 2008). In contrast, we consider that these might be better explained as result of readjustments in the logistic mobility system, due to the functional characteristics of the archaeological sites, with desertion episodes as forest logistic sites whose faunal assemblages are mainly composed of red deer among the ungulates (Botiqueria, Los Banos and Pontet). The potential effects of water availability and vegetation changes on key herbivore biomass, such as red deer, might explain an increase in the risk assumed in hunting parties (Fernândez and Jochim in press). In fact, such readjustments to the logistic mobility system implies the appearance of alternative logistic campsites in high altitude mountain areas specializing in ibex, such as La Cova del Vidre (Bosch 2008), with a clear connection with Lower Aragon. In this sense, one of the clearest examples of the strategic interest of this kind of high-altitude mountain area on land use patterns in the Mediterranean façade of Iberia is represented by the Mas Nou site. This is an open-air site located in the highlands of the limestone plateau of Sierra d'en Seller (Castellon) at 940m a.s.l. with an inhumation pit containing the human remains of seven individuals (Olària et al. 2005). This particular case, which is completely different from the burial contexts associated with coastal or estuarine occupations, could reflect the exertion of property rights over mountain areas as a result of the climatic crisis (Fernândez 2006). The last point to be stressed regarding the impact of the 8200 calBP event on the Iberian Late Mesolithic record relies on taphonomic and archaeological formation processes. In the Mediterranean region of Spain, the presence of a significant stratigraphic hiatus between the Late Paleolithic levels (Magdalenian) and the earliest Neolithic occupations has been recurrently documented in long archaeological sequences in cave deposits such as Cendres Cave, Mallade-tes Cave (Fumanal 1995) and En Pardo Cave (Soler et al. 2008). Even though its genesis regarding such climatic events still needs to be confirmed, it is clear that any Late Mesolithic traces in these archaeological deposits have been beheaded or severely affected. A closer causal relationship might be argued to explain, in the same geographic area, the increasing number of rock shelters, which document chronological gaps, eroded surfaces or the stratigraphic hiatus between the Late Mesolithic (phase A) and the Early Neolithic occupations. At Balma Margineda site (Andorra), level 4S corresponding to a Late Mesoli-thic trapeze phase, was affected by a clear truncation episode which separate this occupations from the Early Cardial Neolithic levels (Berger and Guilaine 2009; Brochier 1995). At Mas de Marti rock shelter (Castellon); several gravel laminations interpreted as eroded surfaces have been recorded between level 3 (Late Mesolithic trapezes phase) and level 2 (Early Neolithic Epicardial phase) (Fernândez et al. 2005). At Tossal de la Roca site, the upper section of the archaeological sequence comprises level I, with a lithic industry of trapezes and a radiocarbon date, overlaid in erosive contact by level Sup, which contains cardial ceramics (Cacho et al. 1995). Finally, at Falguera site, an erosive hiatus 25cm thick (level VII) separates the Late Mesolithic level VIII (with trapezes) from the subsequent Early Neolithic occupations documented in level VI (Garcia and Aura 2006). These examples warn against the problems of detecting the Final Mesolithic in the current archaeological record. Obviously, to avoid biased interpretations, considerations of population dynamics should be grounded on the evaluation of the archaeological evidence on a micro-regional scale, considering the role of site formation processes in the chronological gaps - something that still need to be done. A possible way to detect such terminal Mesolithic evidence in disturbed archaeological contexts is to use individual short-lived samples. For instance, at the Lagrimal Cave in Alicante, the AMS radiocarbon determination of an ibex bone from level IV with anthropogenic fractures and cut marks has provided the most recent chronological evidence for a Late Mesolithic context (Phase B) in the southern Valencia region (Tab. 1). This fact clarifies previous approaches that had outlined the presence of a chronological hiatus of six centuries between the last Late Mesolithic (Phase A) and the first Early Neolithic contexts in this area (Bernabeu 2006). Paleodiets: varying reliance on aquatic resources One of the biggest contributions to Iberian Mesoli-thic archaeological research in recent years has come from paleodiet studies on the basis of ô C13 and ô N15 isotopic analyses. Although this kind of approach can be traced back to the mid-90's in Portugal (Lu-bell et al. 1994), fresh evidence from another Iberian regions such as Cantabria and the Central Mediterranean have recently been published (Arias 2005/ 2006; Garcia Guixé et al. 2006), providing a first data set to compare Late Mesolithic subsistence patterns on a broader scale (Tab. 2). Portugal has the biggest regional data set in terms of number of sites and individuals analyzed (Lubell et al. 1994; Jackes and Meiklejhon 2004; Roksan-dic 2006; [Umbelino 2006in Carvalho 2007]). Notoriously representative, the Muge shell middens record concentrates the highest number of isotopic determinations for the Late Mesolithic period: Cabeço da Arruda (10)5, Moita do Sebastiao (9), Cabeço da Amoreira (5), and Cova da Onça (1). In the Alemtejo 5 The number indicates the quantity of individuals analyzed. region, mainly the Sado estuary, isotopic determinations are less significant quantitatively: Cabeço do Pez (3), Amoreira (2), Arapouco (1), Poças de Sao Bento (1), Vale de Romeiras (1) and Algarao da Gol-dra (2) {Umbelino 2006 in Carvalho 2007). Finally, in the Algarve, the site of Vale de Boi provides a single determination (Carvalho et al. 2008). For the Neolithic period (c. 5200-3000 calBC), the Portuguese data set comprises an outstanding number of sites and determinations, mainly in the Estre-madura region: Gruta do Caldeirao (2), Algar do Pico-to (2), Lapa dos Namorados (1), Lapa da Bougalheira (1), Costa do Pereiro, Algar do Barrao, Lugar do Canto, Casa da Moura (4), Algar do Bom Santo (5), Ped-reira de Salemas (1), Gruta do Correiro Mor (1), Monte do Castelo, and Lapa do Fumo (Carvalho 2007). The first paleodiet studies undertaken for Muge sites estimated a diet contribution of aquatic origin proteins at about 40-50% (Lubell et al. 1994). However, the pattern is more complex if radiocarbon chronology and new isotopic determinations are considered. For instance, Jackes and Meiklejohn (2004) suggest a trend along the Late Mesolithic occupation towards a more terrestrial diet, which they interpret in terms of decreasing shellfish availability due to variations in tidal influence at the Lower Tagus estuary (Van der Schreck et al. 2007). Furthermore, the relationship between changes in the estuarine regime and the reliability of aquatic resources in Muge has been pointed out to interpret the isotopic values from the oldest individual (skeleton 6 from Cabeço da Arruda), which has a lower significant marine protein contribution (24%) and who lived right before the estuary's maximum extent (Martins et al. 2008). Thus, the Muge area points to a sequence of dietary changes closely related to the local evolution of the estuarine regime, which led to the formation of rich shell beds: a first occupation with a lower reliance on aquatic resources (24%), followed by a phase with a significant increase in marine protein input (45-50%) and, finally, a trend towards a more terrestrial diet, when the tidal influence decreased, becoming less productive in terms of shellfish availability. Apart from Muge, the available data from other Portuguese regions during the Late Mesolithic is quite variable. For instance, at the Sado estuary, the shell midden of Arapouco reflects a comparable aquatic protein input to that of the Muge sites, while other Late Mesolithic sites located far inland, such as Amo- reiras and Cabeço do Pez, reflect a more terrestrial input. This dual pattern might be interpreted as the result of the geographic coexistence of human groups with varying reliance on aquatic and terrestrial resources (Umbelino 2006 in Carvalho 2007). In contrast, the Portuguese data for the Early Neolithic (Caldeirao, Correio Mor, Casa Moura) indicates a significantly different subsistence pattern based on terrestrial origin diet (Lubell et al. 1994; Carvalho 2007). Further North, in the Cantabrian region, the paleo-diet studies of Late Mesolithic samples comprise six individual determinations from two different sites: a shell midden at Colomba (1 individual) and a sepulchral cave at Los Canes (five individuals) (Arias 2005/2006). The isotopic value from the Colomba sample indicates a similar diet contribution from aquatic (marine) and terrestrial proteins. This pattern is very similar to those obtained in other, older Cantabrian shell middens dated to the Boreal period, such as J3 and Poza l'Egua. In contrast, the five individual determinations from the Los Canes site show the main protein contribution was of terrestrial origin. On this basis, P. Arias has suggested the coexistence of different populations with a differential reliance on marine and terrestrial resources during the Late Mesolithic. For the Neolithic period, the only Cantabrian determination comes from the Megalithic tomb of Cotero de la Mina, which indicates a terrestrial origin diet (Arias 2005/2006). In the Mediterranean region, the shell midden at El Collado is the only Late Mesolithic site with paleo-diet reconstructions based on stable isotopic analysis. This site includes a cemetery with fifteen individual burials, with nine skeletons analyzed (Garcia Guixé et al. 2006). The most striking result from this shell midden is the low reliance on marine foods -only 25% for two individuals, with the highest ô 13C values, while the rest of the sample has lower values, and in three cases the diet proteins are all terrestrial. Garcia Guixé et al. interpret these results considering two hypotheses: either a different dietary adaptation, where the use of terrestrial resources played a higher role in subsistence patterns than in other European Mesolithic populations, or the less productive nature of Mediterranean sea shellfish compared to Atlantic species. Although the current body of data is still too thin to reach meaningful conclusions on the Iberian scale, several observations can be made. First, a different evolutionary pattern of reliance on aquatic resources can be established in those areas where anthropological series allow a diachronic approach (Portugal and Cantabrian strip). In Portugal, a clear trend towards an increasing contribution of aquatic diets from the Boreal Mesolithic to the Late Mesolithic is observed, which fit with changes documented in the settlement pattern. In contrast, the Cantabrian samples record significant aquatic protein input in the Boreal period, while the Late Meso-lithic ones show the opposite situation, with samples from coastal sites having about 50% of marine proteins and an inland site whose diet is mostly terrestrial. In both regions, Neolithic samples indicate a dietary shift, with a higher contribution from terrestrial proteins. Second, the reliance on aquatic resources between the Muge and the Mediterranean samples is considerably different. It is true that in the El Collado case, the lower number of radiocarbon dates (just two determinations from the same individual) does not allow a determination of diachronic trends. However, it should be noted that the same kind of adaptations - shell middens - led to different results in terms of protein intake. Third, the current data set suggests the coexistence of neighboring coastal and inland Mesolithic populations, with different diets. As Arias states, the Can-tabrian case mentioned above is paradigmatic, but a similar situation has been recently suggested for the Sado estuary (Martins et al. 2008). Finally, there are no discernible differences in isotope values between males and females in any studies published thus far. Even though it is always necessary to incorporate new sites and samples, the main hypothesis is that during the Neolithic transition, the last Iberian hunter-gatherer populations faced various cultural contacts under distinctive subsistence conditions. The dispersal of agriculture in Iberia: chronological framework and spatial distribution In the last seven years, remarkable advances have been achieved regarding the chronology and geography of the initial spread of farming in Iberia. Unlike previous Iberian syntheses (Zilhäo 2000; Bernabeu 2002; Juan-Cabanilles and Marti 2002), mainly focused in traditional research areas such as the Spanish Mediterranean coast and Portugal, new Neolithic archaeological sites located in Iberian Meseta or in the Cantabrian façade have provided fresh evi- dences, including direct radiocarbon dates on crops (Zapata et al. 2004; Stika 2005). Figure 3 and Table 3 show the geographic location and absolute chronology of the earliest farming contexts dated by radiocarbon in the different Iberian regions. The distribution information for all 38 dates is plotted in Figure 4. Considering both the radiometric data and the geographical distribution of the present evidence, some observations on the timing and spatial patterning of first Neolithic contexts in Iberian Peninsula can be made. • The earlier antiquity of the Iberian Neolithic coastal regions (except the North façade) and its clustered and discontinuous spatial distribution. Dates from the Mediterranean coast of Spain (Mas d'Is and Nerja cave) and Portuguese Estremadura remain the oldest Neolithic contexts (5600-5500 calBC), that is, assemblages of pottery and domesticates (plants and animals). The chronology of these first farmer sites SITE CONTEXT REF. LAB. DATA BP S13C (0\00) S15N (0\00) Ref. PORTUGAL Moita do Sebastiao Ossada 22 TO-131 7240±70 -16.1 12.2 1 Moita do Sebastiao Ossada 29 TO-133 7200±70 -16.9 10.4 1 Moita do Sebastiao Ossada 24 TO-132 7180±70 -16.8 11.9 1 Moita do Sebastiao Ossada 41 TO-134 7160±80 -16.7 11.2 1 Moita do Sebastiao Skeleton 16 Beta-127449 7120±70 -16.8 - 2 Moita do Sebastiao Ossada CT TO-135 6810±70 -15.3 13.4 1 Cabeço da Amoreira Skeleton ? TO-11819R 7300±80 -16.3 - 3 Cabeço da Amoreira Skeleton 7 Beta-127450 6850±40 -16.5 11.9 2 Cabeço da Amoreira Burial CAM-00-01 TO-10218 6630±60 -17.1 - 4 Cabeço da Amoreira Burial CAM-01-01 TO-10225 6550±70 -20.1 8.2 4 Cabeço da Arruda Skeleton 6 Beta-127451 7550±100 -19 - 2 Cabeço da Arruda Burial CA-00-02 TO-10216 7040±60 -17.9 10.6 4 Cabeço da Arruda Ossada III TO-360 6990±110 -17.7 11.2 1 Cabeço da Arruda Ossada A TO-354 6970±60 -19 12.2 1 Cabeço da Arruda Ossada 42 TO-359a 6960±70 -17.2 11.8 1 Cabeço da Arruda Ossada D TO-355 6780±80 -18.9 10.3 1 Cabeço da Arruda Burial CA-00-01 TO-10217 6620±60 -18.1 10.5 4 Cabeço da Arruda Ossada N TO-356 6360±80 -15.3 12.5 1 Cova da Onça Skeleton ? Beta-127448 7140±40 -17.2 - 2 Vale Boi c.2 TO-12197 7500±90 -18.3 11.6 5 MEDITERRANEAN REGION AND EBRO VALLEY El Collado Ind. 1 (indet.) - - -19.5 10.2 6 El Collado Ind. 2 (female) - - -19.1 8.9 6 El Collado Ind. 3 (male) - - -17.6 10.2 6 El Collado Ind. 4 (male) - - -17.6 12.8 6 El Collado Ind. 5 (female) - - -18.2 10.6 6 El Collado Ind. 6 (male) - - -18.2 10.9 6 El Collado Ind. 7 (female) - - -17.9 8.9 6 El Collado Ind. 12 (male) UBAR-281 7640 ±120 -19 9.5 6 El Collado Ind. 12 (male) UBAR-280 7570 ±180 - - 6 El Collado Ind. 13 (male) - - -18.1 10.4 6 CANTABRIAN STRIP Colomba Shell midden TO-10223 7090±60 -15.8 12.5 7 Los Canes 6—III AA-6071 6930±05 -19.3 7.7 7 Los Canes 6—II (feet) AA-5295 6860±65 -19.2 9.4 7 Los Canes 6—II (skeleton) AA-5296 6770±65 -19.7 8.1 7 Los Canes 6—II (skeleton) AA-11744 7025±80 -19.6 7.8 7 Los Canes 6-I AA-5294 6265±75 -20 7.9 7 Tab. 2. Stable isotopic values (S13C and S15N) of Late Mesolithic skeletons from different Iberian regions. References: 1: Lubell and Jackes 1994; 2: Umbelino 2006; 3: Detry 2007; 4: Roksandic 2006; 5: Carvalho 2007; 6: Guixé et al. 2006; 7: Arias 2005/2006. partly overlaps. Moreover, their geographical distribution offers a clustered pattern in several regions, such as in southern Valencia, central Catalonia, Algarve and Estremadura, and south of Andalusia. The Valencia region is now the area with the oldest radiocarbon dates on cereals retrieved from open-air sites like Mas d'Is (Bernabeu et al. 2003). Caves were used as habitations (e.g. Or Cave, Cendres Cave) (Marti et al. 1980; Bernabeu et al. 2001), and together with rock shelters, they were also used for stabling (e.g. Falguera) (Garcia and Aura 2006). Recently, a new Early Neolithic site resulting from rescue excavations - El Barranquet - has yielded an old radiocarbon date in sheep bone (Esquembre et al. 2008), and one ceramic context dominated by impressed non-Cardial ceramics very similar to the Li-gurian phasies of Sillon d'Impressions documented in southern France (Bernabeu et al. in press). Central Catalonia is the second cluster of Early Car-dial Neolithic sites documented in Iberia, containing information of both open-air sites (village dwellings) and caves, but until recently poorly dated with radiocarbon determinations based on non-determined charcoal samples (Mestres 1995). Recently, Can Sa-durni Cave has provided the earliest dates on cereal for the area (5470-5300 calBC), although the sample is a cluster of charred cereals (27 gr.) retrieved from the same ceramic ware (Blasco 2005). In Portugal, the empirical regional evidence of some of the earliest Neolithic contexts derive from the western Algarve and northern Estremadura (Zilhäo 2000; Carvalho et al. 2008). Radiocarbon determinations from the early Neolithic contexts in the Algarve region are based on shell samples (Cabranosa and Padrao) (Zilhäo 1997.36). In northern Estremadura, the caves of Almonda and Caldeirao have yielded dates from short-lived materials (adornments, and sheep and human bones) dated to 5400 calBC and 5300 calBC, respectively (Zilhäo 1992; 2001). Finally, in the southern Mediterranean region of Andalusia, Nerja cave contained one of the oldest sheep bone samples (c. 5600-5400 calBC), although the archaeological context of the sample was disturbed (Aura et al. 2005). • Agriculture rapidly reached the deep interior of Iberia. Traditionally, Spanish research considered some Cardial inland enclaves such as the site of Cha-ves and Forcas II in Upper Aragon (Utrilla et al. 1998) and Carigüela in Upper Andalusia (Juan-Ca- Fig. 3. Regional distribution of the earliest farming contexts in different Iberian regions (see Tab. 3 and Fig. 4). SITES CONTEXT SAMPLE MATERIAL SAMPLE REFERENCE 14C age (BP) CalBP 20 range CalBC 20 range Ref. MEDITERRANEAN AREA Mas D'Is 80205 S H. vulgare Beta-166727 6600 ± 50 7613-7335 5664-5386 1 Mas D'Is 80219 S H. vulgare Beta-162092 6600 ± 50 7613-7335 5664-5386 1 Or Basal Cardial (1955-58) S Cereal KN51 6510 ± 160 7855-6886 5906-4937 2 Or J4, level 17 S T. aestivum OxA-10192 6310 ± 70 7430-6960 5481-5011 2 Or J4, level 14 S T. aestivum OxA-10191 6275 ± 70 7425-6937 5476-4988 2 Or Upper Cardial (1955-58) S Cereal H-1754/1208 6265 ± 75 7424-6904 5475-4955 2 Oliva UE79 B Ovis aries Beta-221431 6510 ± 50 7566-7272 5617-5323 3 Falguera 2051b S T. monococcum Beta-142289 6510 ± 80 7609-7174 5660-5225 4 Cendres H16 S T. dicoccum Gif-10136 6490 ± 90 7620-7160 5671-5211 5 Chaves Ib F Acorn GrA-28341 6380 ± 40 7427-7172 5478-5223 6 Chaves Ia H Homo GrA-26912 6230 ± 45 7278-6948 5329-4999 6 Can Sadurnf 18 S cereal UBAR-760 64°5 ± 55 746°-7167 5511-5218 7 Nerja NV-2 B Ovis aries Beta-131577 6590 ± 40 7578-7421 5629-5472 8 PORTUGAL Almonda Level 1 P Cervus elaphus OxA-9287 6445± 45 7421-6851 5529-5305 2 Almonda Level 1 Bb - OxA-9288 6445± 45 7315-6675 5529-5305 2 Caldeirao NA2 B Ovis aries OxA-1035 6330 ± 80 7478-7254 5516-5001 2 Caldeirao NA2 B Ovis aries OxA-1034 6230 ± 80 7478-7254 5472-4902 2 Caldeirao NA2 H Homo OxA-1033 6130 ± 90 7465-6950 5366-4726 2 MESETA La Paleta Pit 219 S Cerealia sp. Beta-223092 6600 ± 60 7622-7322 5673-5373 9 Mirador Level MIR 23 S T. dicoccum Beta-208134 6300 ± 50 7420-7008 5471-5059 10 Mirador Level MIR 22 S T. aestivum\durum Beta-208133 6110 ± 40 7167-6797 5218-4848 10 La Vaquera Level 94 F Acorn GrA-9226 6440 ± 50 7491-7177 5542-5228 11 La Revilla E 9 S Cereal UtC-13347 6313 ± 48 7421-7020 5472-5071 12 La Revilla E 2 S Cereal UtC-13269 6250 ± 50 7323-6 9 49 5374-5000 12 La Revilla E 2 S Cereal UtC-13350 6210 ± 60 7315-6 9 49 5366-5000 12 La Revilla Pit 12 S Cereal UtC-13295 6250 ± 50 7323-6049 5374-5000 12 La Revilla Pit 12 B Ovicaprid KIA-21353 6156 ± 33 7243-6946 5294-49 97 13 La Revilla Pit 4 B Ovicaprid KIA-21356 6355 ± 30 7415-7177 5466-5228 13 La Revilla Pit 4 B Sus sp. KIA-21359 6245 ± 34 7270-7002 5321-5053 13 La Revilla Pit 4 S Cereal UtC-13348 6120 ± 60 7235-6897 5286-4948 13 La Revilla E 13 B Ovicaprid KIA-21354 6177 ± 31 7318-6885 5369-4936 13 La Lampara Pit 1 S T. monococcum UtC-13346 6280 ± 50 7417-6996 5468-5047 13 La Lampara Pit 1 H Homo KIA-6740 6144 ± 46 7247-6856 5298-4907 14 La Lampara Pit 1 B Bos s.p. KIA-21348 6125 ± 33 7167-6882 5218-4933 13 CANTABRIA BASIN Arenaza Ic2 B Bos Taurus OxA-7157 6040 ± 75 7241-6640 5292-4691 15 Miron 3033 S T. dicoccum GX-30910 5550 ± 44 6467-6216 4518-4267 16 Kobaederra level 1 S H. vulgare AA-29110 5375 ± 90 6414-5901 4465-3952 17 Pico Ramos - S H. vulgare Beta-181689 5370 ± 40 6290-5994 4341-4045 17 Tab. 3. Early Neolithic radiocarbon dates on short lived samples. B = bone; Bb = bone bead; H = human bone; F = fruit; P = pierced canine; S = seed. References. 1. Bernabeu et al. 2003; 2. Zilhäo 2001; 3. Esquem-bre et al. 2008; 4. Bernabeu 2006; 5. Bernabeu et al. 2001; 6. Utrilla et al. 2008; 7. Blasco et al. 2005; 8. Aura et al. 2005; 9. Jiménez et al. 2008; 10. Vergés et al. 2008; 10. Estremera 2003; 12. Stika 2005; 13. Rojo et al. 2006; 14. Rojo and Kunst 1999; 15. Arias et al. 1999; 16. Pena-Cocarro et al. 2005; 17. Zapata et al. 2005. banilles and Marti 2002). However, new evidence from the North Meseta show that incursions were not isolated but distributed over a wide area in the hinterland during the same chronological span, c. 5300-5200 calBC. Chaves Cave, a site located in the southern pre-Pyre-nean foothills, contains two Neolithic levels (Ia and Ib) with storage pits and hearth structures. The shortlived dated materials, an acorn from level Ib and a human bone from level Ia, have yielded a slightly younger chronology (5300-5200 calBC) than the previous dates based on charcoal samples (Utrilla et al. 2008). In the North Meseta, the first Neolithic occupations occurred almost coevally in different locations where domesticates appear in villages and stabling caves. La Paleta is an open-air site with over 200 hundred pit structures excavated in different chronologies from the Early Neolithic to the Bronze Age. A preliminary paper has published one radiocarbon date from a cerealia sp. seed, which derives from the organic material used as temper in anthropomorphic ware with cardial decoration; this date is statistically similar to the oldest coastal Iberian Neolithic contexts - 5673-5373 calBC - (Jiménez-Guijarro et al. 2008). Nevertheless, the presence of such an old occupation associated with cardial pottery should be confirmed and better characterized in future publications. In the case of La Vaquera Cave (Valladolid), an acorn recovered in level 94 that belongs to the first and most intense occupation of the cave - Phase I - was dated to the second half of the sixth millennium calBC - 5542-5228 cal BC- (Estremera 2003). Associated with this material, in the same level, other indirect indicators of farming activity such as wheat seeds and blades with use-wear sickle polish have also been found. In the Atapuerca chain, the El Mirador Cave stratigraphy recently produced a detailed sequence spanning from the Early Neolithic to the middle/upper Bronze Age. For the purposes of this paper, the earliest acceptable radiocarbon date comes from two cereal seeds found in the lowest levels of the cave (22 and 23) and dated to the beginning of the Neolithic, i.e. to c. 5400-5200 calBC (Vergès et al. 2008). Finally, in the Ambrona Valley, there is a long series of radiocarbon dates that situate the appearance of the Neolithic at the beginning of the sixth millenni- um calBC. However, considering only short-lived materials unequivocally identified as domesticates (plants or animals), La Revilla produced dates in the last third of the 6th millennium calBC (c. 54725071 to c. 5374-5000 calBC). Again, a single cereal grain (Stika 2005) and one animal bone fragment yielded the oldest Neolithic radiocarbon dates in La Lampara (c. 5400-5200 calBC). Among the archaeological records from both sites, a large number of pit structures have been identified, mainly storage pits in the former, and some pit structures (storage pits, rubbish pits), but also a singular ditched enclosure in the latter. The extended sequence of 14C dates provided by samples coming from the same pit suggests certain problems. Since the same space was continuously reused, materials from different periods have been displaced by post-depositional processes and become part of the filling in some features. That is the case of pit 4 in La Revilla, where three consecutive radiocarbon dates (6355+30 BP, 6245+34 BP, and 6120+60 BP), separated by some centuries when calibrated (Tab. 3), appeared at different depths in a disorderly stratigraphic sequence. Subsequently, the interpretation of the occupation phases of these sites remains controversial, above all regarding the degree of mobility developed by early farmers. • The later appearance of agriculture and domesticates in the Cantabrian façade. Direct radiocarbon evidence of domesticates in this region displays a significant chronological delay compared to other areas such as North Meseta or the Mediterranean and Portuguese coasts. The chronological frame is variable, depending on whether the radiocarbon samples are from domestic cattle or crops. On the one hand, Are-naza Cave has a 'post-Mesolithic' level (Ic = IC2) with geometric microliths, undecorated ceramics and two bovine remains classified as cattle (Bos taurus), the oldest being directly dated to c. 5292-4691 calBC (Arias and Altuna 1999). Nevertheless, a third cattle jaw bone, supposedly from the same level, yielded an AMS date of 10 860+120 BP. Thus, it seems that the stratigraphy is plagued by uncertainties. On the other hand, the oldest cereal in the Cantabrian area, from El Miron Cave, has been dated to the second half of the fifth millennium calBC; level 303.3 provided a charred grain identified as Triticum dicoccum (emmer), dated to c. 4518-4267 calBC (Pena-Cho-carro et al. 2005). Lastly, we can mention two dates from barley grains, one associated with a few high-quality combed and digitally impressed ceramics, geometric microliths and marine mollusks in Kobae-derra cave (c. 4465-3952 calBC) (Zapata 2002); and OxCal v4,1.3 Bronk Ramsey (2009); r:5 lntCal04 atmospheric curve (Reimer et al 2004) 6500 6000 5500 5000 4500 4000 3500 Mas d'ls (6600±50BP) Mas d'ls (6600±50BP) Or (6510±160BP) Or (6310±70BP) Or (6275±70BP) Or (6265±75BP) Oliva (6510±50BP) Falguera (6510±80BP) Cendres (6490±90BP) Chaves (6380±40BP) Chaves (6230±45BP) Can Sadurni (6405±55BP; Nerja (6590±40BP) ( = = c » = 1 0= ( « [ = r ■= c«~ ■I 0 Hordei. ■i 0 Hordet > Ovis ar =■■=■ in Tritic =■-J Tritic (iisX) Aco ™ ZD Ce 3 Ovis a ries Cereal ;es um monoccoci um diccocum n Homo eai MEDITEf stivum stivum ' m !RANEANAREA Almonda (6445±45BP) Almonda (6445+45BP) Caldeirâo (6330±80BP) Caldeirâo (6230+80BP) Caldeirâo (6130±90BP) r t 1 — ) Cervi .= 1 Ronfi a < > L • s elaphus bead Ovis a ries 1 Ovis aries 1= —> Hnii 10 PORTUGAL La Paleta (6600±60BP) Mirador (6300±50BP) Mirador (6110±40BP) La Vaquera (6440±50BP) La Revilla (6313±48BP) La Revilla (6250±50BP) La Revilla (6210±60BP) La Revilta (6250±50BP) La Revilla (6156±33BP) La Revilla (6355±30BP) La Revilla (6245+34BP) La Revilla (6120±60BP) La Revilla (6177±31BP) La Lampara (6280±50BP) La Lampara (6144±46BP) La Lampara (6125+33BP) c° = r :—) Cereali, C =>v—1 1= = —= 1 D Ar.nri r =11—1 r = ». c® ■=■= t = ... ( =• 1 Oi/f. OCi.= Of ) rs . J sp. Triticum diccoc 1= JO Triticum Cereal Cereal 1 Cereal Cereal J Ovis/Capra i/Capra Sus domestic " j Cereal Ovis/Capra Triticum mon( 1 Homo Bos sp. / um aestivum JS 1 ccocum JORTH MESETA Arenaza (6040±75BP) El Miron (5550±40BP) Kobaederra (5375±90BP) Pico Ramos (5370±40BP) 1 ■ • ,..:..:!.:....... m— « He CANTABRh = —> Bo ( rdeum vulgare \N FAÇADE s taurus Tritini. m diccocum i Hordeum vulgare 6500 6000 5500 5000 4500 4000 3500 Calibrated date (calBC) Fig. 4. Chronological framework of farming dispersal in Iberia, with the earliest radiocarbon dates on short lived sample (see Fig. 3). the second one at the Pico Ramos site (c. 4341-4045 calBC), a small cave where cereal was found, but neither ceramics nor domesticated animals (Zapata 1995). Revisting the distribution of farming on the Iberian peninsula: discussion In the light of new paleo-environmental information and archaeological data on the Late Mesolithic record, and taking into consideration current chronological information and the distribution pattern of the earliest Neolithic contexts on the Iberian Peninsula, we can raise new questions about how farming spread during the Mesolithic-Neolithic transition. Regarding previous syntheses (Zilhäo, 2001; Juan-Cabanilles and Marti 2002), the main question still turns on discovering the driving forces behind the rapid Neolithic expansion through the interior Iberian regions. In about 200 years (or nearly 8 generations), all the components of the 'Neolithic package' (domesticates, pottery, village dwellings, houses) simultaneously appear from the oldest coastal enclaves and throughout the interior locations. This phenomenon requires a reformulation of traditional expansion mechanisms previously invoked to explain the early appearance of the Neolithic in the Spanish Meseta (Rojo et al. 2006.85-86). Furthermore, a more complex historical rather than contingent (migration vs. adoption) perspective is need to consider both the demographic and cultural factors behind the underlying migration mechanisms. Bearing in mind the updated chronological overview presented above and the most recent archaeological evidence, we can inquire into the validity of the main farming dispersal models proposed for the Iberian case so far: the cultural diffusion model, maritime pioneer colonization and demic diffusion. In Iberia, the most elaborated cultural diffusion model is the 'Capillary Model' (Vicent 1997; Cruz and Vicent 2007; among others). This model assumes continuity of human occupation between the Meso-lithic and Neolithic periods to explain the introduction and spread of pottery and domesticates in hunter-gatherer social networks. Rather than refute its validity here, we wish simply point out that this modality could be considered only in those areas where archaeological evidence of local hunter-gatherers exists; in other words, where Late Mesolithic sites would have been documented. In this sense, this model is not adequate to explain Neolithic distribution in the North Meseta or in Catalonia because of the lack of Late Mesolithic sites in these regions, as shown above. However, it also should be noted that some of the theoretical assumptions of this hypothesis, such as seasonal storage and cycles of delayed return consumption (Vicent 1997) are completely absent from the Iberian Late Mesolithic record. On the other hand, this model requires the documentation of transitional forms, for instance the use of pottery or domesticates in Late Mesolithic contexts, to be correctly contrasted. Still, the identification of this situation in the archaeological record is complex, given the necessary critical evaluation of archaeological formation processes that lead to these kinds of association (Zilhäo 1998; Bernabeu 2002; 2006; Juan-Cabanilles and Marti 2006). At present, most of the Iberian Late Mesolithic contexts with ceramics are documented in rock-shelters consisting of multi-stratified archaeological deposits with ceramic and preceramic levels represented in the stratigraphy, or even in the same stratigraphic unit. In all these cases, the relationship between cultural remains and the dated material should be regarded with caution, given the taphonomic factors, the sedimentary processes and potential post-depositional disturbance6. The Cantabrian coast is the only Iberian region where adoption processes were likely to have occurred, given the long geographic coexistence of foraging groups there with the neighboring North Meseta and High Ebro Valley farming populations during the last 6 During last years, the most debated study case of ceramics found in terminal Mesolithic contexts is represented by Mendandia site, where the oldest ceramics were retrieved from Level III-Sup dated in 7210±80 BP and 7180±45 BP. As A. Alday has pointed out, the typology of the ceramic assemblage of this level is simpler (mainly bowls and 'S' shaped wares) and stylistically different (mainly slight incisions and some plain cords) than the Mediterranean Impressed and Cardial complexes. This author has defended the validity of such chronology invoking for a some kind of parallelism with other Western European Mesolithic cultures, which developed their own ceramic styles such as the Rocadourien, La Hoguette and Limbourg complexes (Alday 2005.489). However, even accepting the fact that Mendandia might represent a kind of different ceramic style, there are stratigraphic and archaeological arguments to refuse such old chronology: First, level III whose stratigraphic depth is about 25 cm was arbitrarily divided in two sections, the oldest (III inf) preceramic and the more recent ceramic (III sup). Regarding to this, the nature of radiocarbon samples from level III sup are clusters of bones (250 gr from the excavation unit Z2 and 115 gr from Z3) retrieved at different depths (-120 to -125 cm from excavation unit Z2 and -125 to -130 from excavation unit Z3). This fact implies that each radiocarbon date is composed by aggregates of organic materials from different occupation episodes that were grouped and interpreted as singular occupational events; Second, the chronology of level III sup falls into the chronological limits of Late Mesolithic phase A and quarter of the sixth and the first half of the fifth millennia calBC. A second farming dispersal model proposed for the Iberian case is Maritime Pioneer Colonization (MPC) (Zilhäo 1993; 1997; 2001). According to the most recent radiocarbon evidence, the appearance of the Neolithic in Iberia still fits with the expectations of MPC, as proposed by Zilhao. The earliest data on farming are linked to coastal or neighboring coastal areas on the Mediterranean coast of Spain (Central Catalonia, South Valencia and Malaga) and Portugal (the Algarve and North Estremadura). No evidence or very little evidence of previous Late Mesolithic occupations is known in those areas, suggesting an elusive, non-overlapping strategy for the first Neolithic installations regarding local forager populations. Expansion by sea, which produced long-distance relocation episodes in early Neolithic enclaves, could potentially explain the rapidity of expansion, as well as similarities in material culture (especially in ceramic decoration). However, the main interpretative challenge for this model still comes from the driving forces that caused this mode of farming dispersal. In this sense, when comparing several Iberian Neolithic enclaves, the archaeological record contains significant differences in terms of material culture (suggesting different cultural affiliations) and settlement data (site densities, rock art, habitat features, etc). Moreover, the cultural homogeneity of the first ceramic contexts is the subject of recent discussion. Recently, Manen et al. (2007) and Carvalho (in press) have suggested a possible Maroquain origin for the Early Neolithic in South Portugal on the basis of some similarities in ceramics (back-lip shapes and almagra) and lithics (heat treatment, pressure debi-tage, and dominance of lunates among the micro-liths). As noted by these authors, these characteristics are not common in traditional Cardial Franco-Iberian archaeological culture. In addition, Bernabeu (Bernabeu et al. in press) has recently identified an early Neolithic context at the El Barranquet site (Oliva, Valencia) dominated by impressed non-Cardial ceramics similar to the Ligu- rian phacies of Sillon d'Impressions documented in southern France (sites of Peiro Signado and Pont de Roque-Haut) and northern Italy (Arene Candide). Bernabeu goes further suggesting that this ceramic impressed horizon could even be pre-cardial, representing the first Neolithic cultural complex in Iberia before 5500 calBC, and being on the bases of subsequent Cardial culture and the earliest North Meseta Neolithic contexts (i.e. La Revilla, La Lampara sites). On the other hand, it should be noted that the archaeological record indicates significant differences in complexity (number of sites, density of occupation, rock art and monuments) among the different early Neolithic enclaves. For example, the southern Valencia region displays an Early Neolithic Cardial record of villages, rock art sanctuaries, and monumental enclosures (Bernabeu et al. 2003) significantly much more complex than sites in the Algarve or Portuguese Estremadura regions. To recapitulate, as regards the first farming enclaves, a developing variety and complexity not contemplated by the MPC model assumptions, has been discovered. Placed in the general Iberia context, such differences in complexity and settlement densities might indicate the existence of social presure and attractive factors such as segregation and aggregation processes governed by rank and social competition strategies (Anthony 1997; Fiedel and Anthony 2003). Finally, demic diffusion has also been invoked as an interpretative model to explain the populating of many Iberian regions during the Early Neolithic Epi-cardial phase (c. 5200-4800 calBC) (Juan-Cabanilles and Marti 2002; Carvalho in press). The chronological framework and the regional distribution of sites of this chronology attest to a subsequent and gradual expansion process from the Cardial enclaves through neighboring or adjacent areas following fission and short-distance relocation logic (e.g. Mestres 1992). However, the earliest radiocarbon dates provided by the North Meseta sites (c. 5400-5300 calBC) cannot be simply explained by this model, because of the geographic discontinuity and distance from the coastal Neolithic enclaves (more than 500km). On the about 6 centuries older than the first Cardial or Impressed ceramic contexts found in the Iberian coasts dated on the basis of strati-graphic associations with short lived individual samples (see Fig. 4). However the lithic industry, dominated by lunates with bifacial retouch among the geometric microliths is discordant too with the chronology displayed by this type of armatures in Iberia (see Fernândez et al. 2009). Finally, considering a micro-regional scale validation, the appearance of ceramics at Mendandia is about 700 years older than other neighbor sites such as Zatoya, Atxoste or Aizpea where the first ceramics are found overlying Late Mesolithic contexts (Alday 2003). As J. Bernabeu has recently stated (2006) it is difficult to accept Mendandia as the only site in the High Ebro Valley where pottery is present during 600-800 years while in other neighbor locations is absent. one hand, this fact implies a lower rate of spread than MPC (10-20km per year), but higher than demie diffusion (1km per year). In this sense, although the present paleo-anthropological data on the Iberian Peninsula is too limited to provide enough information to test the demographic growth of farming communities (also known as the Neolithic demographic transition) recognized in other European regions (Bocquet-Appel 2002; 2008), population growth does not seem to have been the only cause for the rapid dispersal that the radiocarbon dates indicate. Moreover, as Shennan remarks, in Central Europe (Shen-nan 2007) demographic growth would not necessarily have triggered spatial expansion, since settlement data show that new places were colonized before others reached any sort of maximum carrying capacity. Additionally, this situation would demand alternative farming dispersal models to set off the migration of coastal farming groups to inland regions. For instance, long distance inland colonization as a result of fission and settlement relocation logic driven by household decision-making might be invoked (Bogu-cki 2003; Marti 2008). Similarly, an ideal free distribution pattern as a spatial behavior might have been carried out by farmers, whereby they tended to occupy sites giving the best yields (Shennan 2007; McClure et al. 2009). According to this view, in the very first wave of migration, household decision-making would have consciously evaluated the costs and benefits of farming strategies and settlement relocation according both to the ecological suitability of the territories and dynamic social networks (Boguki 2003). In this sense, the spectrum of crops identified at Early Neolithic sites in the northern Meseta and the agrarian practices involved, indicate agricultural activity was well-established and perfectly adapted to the local ecological conditions (Zapata et al. 2004). It is likely that crops identified in La Lampara and La Revilla sites, hulled wheats and barley, might have been selected because they were suited to that area (resistance to poor soil conditions and fungal disease; toleration of drier conditions). The last issue to be discussed here is how the distribution pattern of both Late Mesolithic and Early Neolithic settlements and the chronological trends in the spread of farming would be reflected in the genetic composition of ancient Iberian populations. The key question, however, is whether the farming transition was a process of cultural and economic change, but not marked gene flow, or if it implied the arrival of genetically distinct populations which replaced the Mesolithic ones; this matter should be addressed in terms of the regional variability of socio-cultural processes. During the 90's, anthropological studies reached contradictory conclusions on this issue. On the basis of skeletal differences Lalueza-Fox (1996) suggested population replacement, while others have observed that there is no evidence of discontinuity in Portuguese dental morphology (Jackes et al. 1997; 2001). In the same way, using craniometric data in a broader European context, Pinhasi and Pluciennik see no marked difference between Late Mesolithic and Early Neolithic human remains in the Western Mediterranean region, which might be interpreted as a result of a higher level of biological admixture (Pinhasi and Pluciennik 2004, but see also Zilhao 2004 for comments on biases in the Iberian data set). From the beginning of the 21st century, this matter has been addressed on the basis of molecular studies of archaeological data. However, in dealing with the genetic issues involved in the Iberian Peninsula Mesolithic-Neolithic transition, two major problems arise: O There is remarkable regional disparity in the relevant data sets to address this question, most of them from Portugal (Bamforth et al. 2003; Chandler et al. 2005). Obviously, such regional bias challenges any attempt to reach equivalent conclusions for other Iberian areas whose geographic, environmental and demographic conditions were different. Regarding ancient mtDNA, new sites have been analyzed recently in the Mediterranean region, such as Sant Pau (Catalonia) (Fernândez-Dominguez 2005) and Can Grau (Catalonia) (Sampietro et al. 2007); however the skeletal samples do not date to the Neolithic transition, but to the Post-cardial and Middle Neolithic phases, or to the Chalcolithic period (Gamba et al. 2008). © On the other hand, there is no consensus among geneticists regarding the origin of the first European farmers or what one may infer from the geographic distribution of various genetic markers (e.g. Haak et al. 2005 vs. Ammerman et al. 2006 for Central Europe). For instance, Haplogroup J, widely considered one of the main genetic signatures of Neolithic expansions (Wells 2007), is completely absent among the oldest Iberian Neolithic samples, but also lacks for skeletal material analyzed in Syria (Tell Halula and Tell Ramad) dated to the PPNB period (Fernân-dez-Dominguez 2005). Consequently, as far as the Iberian material is concerned, the absence of haplo-group J among the Neolithic samples does not necessarily mean the absence of Neolithic expansion. Bearing in mind the problems mentioned above, the first study published on ancient mtDNA Iberian material, comparing 15 Mesolithic with 13 Neolithic samples, suggests genetic continuity in the female line, given the presence of two individuals from Ar-ruda (Late Mesolithic) and another from Feteira (Neolithic) in haplogroup K (Bamforth et al. 2003). Another interesting conclusion of this study was the lack of haplogroup I (predominant in Northern and Eastern Europe) and U6 (of North African origin). In contrast, a more recent study (Chandler et al. 2005) reports genetic discontinuity between the Early Neolithic (the Algar do Bom Santo, Caldeirao and Perdigoes sites) and the Late Mesolithic populations (the Cabeço da Amoreira, Cabeço do Pez, Poças de Sao Bento, Toledo, Fiais, Vale de Romeiras sites) based on significantly different frequencies of common haplogroups between them; however, their relationship is closer to other Iberian than Near Eastern populations. The authors interpret this discontinuity as support for the MPC model. In this sense, the absence of haplogroup J in the Portuguese Neolithic samples was interpreted as a result of the lack contribution of women of Near Eastern origin in the colonization process, considering the uniparental transmission mode (Zilhäo 2004.78). Concluding remarks and directions for future research It was not our intention in this paper to discuss the reasons behind the adoption of farming, or to discuss in detail the broad array of archaeological situations led by the cultural interaction between foragers and farmers in Iberia. We wish simply to stress some structural problems not widely discussed on the Iberian scale, drawing attention to paleo-environmental evolution and population dynamics. According to the chronological evidence, with the exception of the Northern façade, Iberia witnessed a rapid Neolithic transition process. From the standpoint of evolutionary theories, the spread of farming in Iberia is an example of dispersal opportunities. Early and Middle Holocene environmental changes seem to have affected not only the regional distribution of hunter-gatherer populations, but also the out- comes of such adaptations in terms of subsistence and demography. We have already mentioned that such divergences might have been especially important because of the relationship between coastal and estuarine adaptations, with year-round sedentary or semi-sedentary settlements and potential for demographic growth. For instance, assuming the El Col-lado data as representative of the central Iberian Mediterranean coast, the last foragers relied less on aquatic resources than their Portuguese counterparts in the Tagus estuary, and were possibly more mobile. Consequently, such a structural constrain might explain the rapid extinction of forager subsistence systems after a short period of cultural contact with the first farming groups. Thus, farming expanded to other regions, bringing about different migration processes and interaction phenomena. Even though some regional bias in the Late Mesolithic record needs to be corrected (Andalusia), the current evidence indicates broad areas such as Catalonia and North Meseta had no forager population when the first farmers arrived. In contrast, in those areas where the Late Mesolithic record is present, one could expect a broad range of different cultural contact situations, including small-scale migratory phenomena such as infiltration (Zvelebil 2000), as proposed for the Ebro Valley (Berna-beu 2002). However, the role of forager populations in farming expansion, whether by means of rapid adoption through exchange or assimilation by direct interaction with farming groups, is not clear in the Ebro Valley archaeological record, concentrating current debates. In the same way, it is unwise to neglect a priori the possible contribution of such populations to farming expansion in the North Meseta. It is true that many of the migration debates in this paper center on the North Meseta data. Of course, this region raises new questions about the timing of, and specific ways in which the Neolithic reached the deep interior of Iberia, as a counterpart to other traditional research regions. What seems clear, however, is the need to incorporate new models that present farming dispersal migration as a more complex process driven by social factors, and not merely the simple result of aimless demic growth. -ACKNOWLEDGEMENTS- This research was undertaken as a post-doctoral research project by one of the authors (Javier Fernändez Lopez de Pablo): 'Nuevas técnicas de investigacion y anälisis espacial en arqueologta: la aplicacion de sistemas de information geogräfca y teledeteccion al registro Holoceno de la Comunidad Valenciana', supported by the Direc-cion General de Innovacion Tecnologica Conselleria de Industria y Empresa, Generalitat Valenciana. The radiocarbon date from Lagrimal Cave was obtained in the research project 'Cronolog^a absoluta por Carbono 14 del Epipaleotitco y el Neolüico antiguo en Villena'funded by Fundacion Municipal José Maria Soler of Villena (Alicante) in 2007. We gratefully acknowledge Dr. Joan Bernabeu and Dr. Antonio F. Carvalho for giving us access to unpublished papers, and Dr. Mihael Budja and Dr. Bernat Marti whose suggestions contributed significantly to the final version of this manuscript, none of whom should be held responsible for any views expressed in this paper. REFERENCES ALDAY A. 2002. Las unidades industriales mesoliticas en la alta-media Cuenca del Ebro. Complutum 13:19-50. 2003. Ceramica neolitica de la region vasco-riojana: base documental y cronolögica. Trabajos de Prehisto-ria 60 (1): 53-80. 2005. The transition between the last hunter-gatherers and the first farmers in Southwestern Europe: The basque perspective. Journal of Anthropological Research 61 (4): 469-494. 2006a. El Mesolitico de Muescas y Denticulados en la Cuenca del Ebro y el litoral Mediterrraneo Peninsular: Sintesis de los datos. In A. Alday (coord.), El mesotitico de muescas y denticulados en la Cuenca del Ebro y el litoral mediterräneo peninsular. Diputacion Foral de Alava, Departamento de Cultura: 303-317. (ed.) 2006b. El campamento prehistorico de Mendan-dia: ocupaciones mesotiticas y calcotiticas entre el 8500y el 6500 B.P. Ataun: José Miguel de Barandiaran Fundazioa. Vitoria-Gasteiz: Diputacion Foral de Alava. ALEGRE I. 2005. La industria geométrica en el valle de Am-brona. In P. Arias, R. Ontanön, C. Garcia-Moncö (eds.), Actas del III Congreso sobre el Neolüico en la Penmsu-la Ibérica. Instituto Internacional de Investigaciones Pre-histöricas de Cantabria, Santander: 233-245. ALLUÉ E., ANGELUCCI D. E., CÄCERES I., FIOCCHI C., FON-TANALS M., GARCIA M., HUGUET R., OLLÉ A., SALADIÉ P., VERGÈS J. M., ZARAGOZA J. 2000. El registro paleoecolö-gico y arqueolögico de la Cativera (El Catllar, Tarragona): datos preliminares sobre el limite Pleistoceno-Holoceno en el sur de Cataluna. In V. Oliverira (coord.) 3° Congres-so de Arqueolog^a Peninsular, Vol IX. UTAD, Vila Real: 81-98. AMMERMAN A. J., BIAGI P. 2003. The Widening Harvest. The Neolithic transition in Europe: looking back, looking forward. Archaeological Institute of America. Boston. AMMERMAN A. J., PINHASI R., BÄNFFY E. 2006. Comments on: HAAK et al. - Ancient DNA from the First European Farmers in 7500-Year-Old Neolithic Sites. Science 312: 1875a. ANTHONY D. W. 1997. Prehistoric migration as social process. In J. Chapman, H. Hamerow (eds.), Migrations and Invasions in Archaeological Explanation. British Archaeological Reports S664. Oxford: 21-32. APARICIO J. 1990. Yacimientos arqueolögicos y evoluciön de la Costa Valenciana durante la Prehistoria. II Seminari sobre el Mediterrani «El Mare Nostrum». Academia de Cul-tura Valenciana. Valencia: 7-91. ARAUJO A. C. 2003. O Mesolitico inicial da Estremadura. In V. S. Gonçalves (ed.), Muitagente, poucas antas?: ori-gens, espaços e contextos do Megalitismo. Actas do II Co-löquio Internacional sobre Megalitismo. Instituto Portu-gues de Arqueologia. Lisboa: 101-114. ARIAS P. 1999. The origins of the Neolithic along the Atlantic coast of continental Europe. Journal of World Prehistory 13: 403-464. 2005. Determinaciones de isötopos estables en restos humanos de la regiön Cantabrica. Aportaciön al estudio de la dieta de las poblaciones del Mesolitico y el Neo-litico. Munibe 57:359-374. ARIAS P., ALTUNA J. 1999. Nuevas dataciones absolutas para el Neolitico de la cueva de Arenaza (Bizkaia). Munibe 51:161-171. AURA J. E., BADAL E., GARCIA BORJA P., GARCIA PUCHOL O., JORDÄ J. F., PASCUAL J. L., PÉREZ G., PÉREZ M. 2005. La Cueva de Nerja (Malaga). Los niveles neoliticos de la Sala del Vestibulo. In P. Arias, R. Ontanön, C. Garcia-Mon-cö (eds.), Actas del III Congreso sobre el Neolüico en la Penmsula Ibérica. Instituto Internacional de Investigacio-nes Prehistöricas de Cantabria, Santander: 975-987. AURA J. E., MARCO Y., GARCIA O., JARDON P., JORDÀ J. F., MOLINA L., MORALES J. V., PASCUAL J. L., PÉREZ G., PEREZ M., RODRIGO M. J., VERDASCO C. 2006. Epipaleoiï-tico-Mesolitico en las comarcas centrales valencianas. In A. Alday (ed.), El mesolttico de muescas y denticulados en la Cuenca del Ebro y el litoral mediterrâneo peninsular. Diputacion Foral de Alava, Departamento de Cultura: 65118. AURA J. E., FERNANDEZ LOPEZ DE PABLO J., GARCIA O., JUAN-CABANILLES J., MARTI B. in press. El Mesolitico Geo-métrico de tipo "Cocina" en el Pais Valenciano. In P. Utril-la and L. Montes (ed.), IIReunion sobre Mesolüico de la Cuenca del Ebro y Litoral mediterrâneo. BALBIN R., ALDECOA A., CASADO A B., JIMENEZ P. J., ANDRES B., ALCOLEA J. J., GARCIA M. A. 1997. Avance al estu-dio del poblamiento paleolitico del alto valle del Sorbe (Muriel, Guadalajara). In P. Bueno and R. Balbin (ed.), II Congreso de Arqueologia Peninsular: Zamora, Vol. 1: 201218. BAMFORTH F., JACKES M., LUBELL D. 2003. Mesolithic-Neolithic population relationships in Portugal: the evidence from ancient mitochondrial DNA. In L. Larsson, H. Kindgren, K. Knutsson, D. Leoffler and A. Âkerlund (eds.), Mesolithic on the Move: Proceedings of the 6th International Conference on the Mesolithic in Europe, Stockholm 2000. Oxbow Books. Oxford: 581-587. BARANDIARAN I., CAVA A. 2000. A proposito de unas fe-chas del Bajo Aragon: reflexiones sobre el Mesolitico y el Neolitico en la Cuenca del Ebro. SPAL 9: 293-326. 2001. Cazadores-recolectores en el Pirineo navarro. El sitio de Aizpea entre 8000 y 6000 anos antes de ahora. Anejos de Veleia, Series Major no. 10. Vitoria. BARTON M., RUBIO F., MIKSICEK C. A., DONAHUE D. J. 1990. Domestic Olive. Nature 346:518-519. BERGER J. F., GUILAINE J. 2009. The 8200 calBP abrupt environmental change and the Neolithic transition: A Mediterranean perspective. Quaternary International 200: 31-49. BERNABEU J. 2002. The social and symbolic context of neolithisation. In E. Badal, J. Bernabeu and B. Marti (eds.), Neolithic landscapes of the Mediterranean. Saguntum-PLAV Extra-5: 209-233. 2006. Una vision actual sobre el origen y difusion del Neolitico en la Peninsula Ibérica. In O. Garcia and J. E. Aura (coords.), El Abric de la Falguera (Alcoi, Ala-cant). 8000 anos de ocupacion humana en la cabe-cera del no de Alcoi. Alicante: 189-211. BERNABEU J., PÉREZ RIPOLL M., MARTINEZ VALLE R. 1999. Huesos, Neolitizacion y Contextos Arqueologicos Aparentes. In J. Bernabeu and T. Orozco (eds.), Actas del II Congrès del Neolttic a la Penmsula Ibérica. Saguntum-PLAV Extra 2: 589-596. BERNABEU J., FUMANAL M. P., BADAL E. 2001. La Cova de les Cendres. Volumen 1. Paleogeografa y estratigra-fia. Estudis Neolttics, 1. Universitat de València. Valencia. BERNABEU J., OROZCO T., DIEZ A., GOMEZ M., MOLINA F. J. 2003. Mas d'Is (Penàguila, Alicante): Aldeas y recin-tos monumentales del Neolitico Inicial en el valle del Ser-pis. Trabajos de Prehistoria 60 (2): 39-59. BERNABEU J., MOLINA L., ESQUEMBRE M. A., ORTEGA J. R., BORONAT J. D. in press. La cerâmica impresa mediter-rânea en el or^gen del Neolüico de la Penmsula Ibérica. In Homenage à Jean Guilaine. BICHO N. F. 1994. The end of the Paleolithic and the Mesolithic in Portugal. Current Anthropology 35(5): 664674. BIAGI P. 2003. A review of the Late Mesolithic in Italy and its implications for the Neolithic Transition. In A. J. Am-merman and P. Biagi (eds.), The Widening Harvest. The Neolithic Transition in Europe: Looking Back, Looking Forward. Archaeological Institute of America. Boston: 133-155. BINDER D. 2000. Mesolithic and Neolithic interaction in Southern France and Northern Italy: new data and current hypothesis. In T. D. Price (ed.), Europe's First Farmers. Cambridge University Press, Cambridge: 117-143. BLASCO A. 2005. Primeros datos sobre la utilization sepulcral de la Cueva de Can Sadurni (Bergues, Baix Llobre-gat) en el Neolitico Cardial. In P. Arias, R. Ontanon, C. Garcia-Monco (eds.), Actas del III Congreso sobre el Neolüico en la Penmsula Ibérica. Instituto Internacional de Investigaciones Prehistoricas de Cantabria, Santander: 625-634. BOGUCKI P. 2003. The Neolithic Settlement of riverine interior Europe as a Complex Adaptative System. In R. A. Bentley, D. G. Maschner (eds.), Complex System and Archaeology. Empirical and theoretical applications. University of Utah Press, Salt Lake City: 93-102. BOCQUET-APPEL J. P. 2002. Paleoanthropological traces of Neolithic Demographic Transition. Current Anthropology 43: 638-650. 2008. Explaining the Neolithic Demographic Transition. In J. P. Bocquet-Appel and O. Bar-Yosef (eds.), The Neolithic Demographic Transition and its Consequences. Springer: 35-55. BOSCH A. 2008. La importància de la muntanya en el pro-cés de neolititzacio de la regio del curs inferior de l'Ebre, segons investigacions al jaciment arqueolögic de la Cova del Vidre. In M. A. Pradilla (ed.), Patrimoni i economia a les comarques de la diocesi de Tortosa. Benicarlo: Onada edicions: 49-58. BROCHIER J. E. 1995. Estudi geoarqueologic dels deposits holocens de la Balma de la Margineda, capes de 1 a la 6. In J. Guilaine, M. Martzluff (eds.), Les excavacions a la Balma de la Margineda (1979-1991). Tome 1. Ministeri d'Afers Socials i Cultura. Principat d'Andorra, La Vella: 56-90. BRONK RAMSEY C. 2009. Bayesian analysis of radiocarbon dates. Radiocarbon 51(1): 337-360. BUDJA M. 2007. The 8200 calBP ' climate event' and the process of neolithisation in south-eastern Europe. In M. Budja (ed.), 14th Neolithic Studies. Documenta Praehistorica 34:191-201. CACHO C., FUMANAL M. P., LOPEZ P., LOPEZ J. A., PÉREZ M., MARTINEZ R., UZQUIANO P., ARNANZ A., SANCHEZ A., SEVILLA P., MORALES A., ROSELLO E., GARRALDA M. D., GARCIA M. 1995. El Tossal de la Roca (Vall d'Alcala, Alicante). Recerques del Museu d'Alcoi 4:11-101. CACHO I., GRIMALT J. O., PELEJERO C., CANALS M., SIER-RO F. J., FLORES J. A., SHACKLETON N. 1999. Dansgaard-Oeschger and Heinrich event imprints in Alboran Sea pa-leotemperatures. Paleoceanography 14 (6): 698-705. CARVALHO A. F. in press. Le passage vers l'Atlantique: Le processus de Néolithisation en Algarve (Sud du Portugal). L'Anthropologie. in press. O Mesolitico Final em Portugal. In P. Utrilla and L. Montes (eds.), II Reunion sobre Mesolitico de la Cuenca del Ebro y Litoral mediterräneo. in press. Chronology and geography of the Mesolithic-Neolithic transition in Portugal. In Honoria P. Kalp. 2002. Current perspectives on the transition from the Mesolithic to the Neolithic in Portugal. In E. Badal, J. Bernabeu and B. Marti (eds.), Neolithic landscapes of the Mediterranean. Saguntum-PLAV Extra-5: 135-250. 2003. A emergência do Neolitico no actual territorio português: pressupostos teoricos, modelos interpreta-tivos e a evidência empirica. O Arqueologo Português 21: 65-150. 2007. Novos dados sobre dois temas da Pré-Historia do Sul de Portugal: o Mirense e o processo de neoliti-zaçao. Promontoria 5:91-110. CARVALHO A. F., BICHO N. F., STINER M. C., GIBAJA J. F., VALENTE M. J., MASUCCI M. A. 2005. O projecto "o pro- cesso de neolitizaçao no Algarve" (Portugal): âmbito e pri-meiros resultados. In P. Arias, R. Ontanon, C. Garcia-Mon-co (eds.), Actas del III Congreso sobre el Neolitico en la Peninsula Ibérica. Instituto Internacional de Investiga-ciones Prehistoricas de Cantabria, Santander: 965-974. CARVALHO A. F., DEAN R. M., BICHO N. F., FIGUEIRAL I., PETCHEY F., DAVIS S. J. M., JACKES M., LUBELL D., BEU-KENS R., MORALES A., ROSELLO E. 2008. O Neolitico anti-go de Vale Boi (Algarve, Portugal): Primeiros resultados. In M. Hernandez, J. Soler and J. A. Lopez (eds.), Actas del IV Congreso del Neolitico Peninsular (vol. I). MARQ. Di-putacion de Alicante, Alacant: 267-274. CAVA A. 2000. La industria litica del neolitico de Chaves (Huesca). Salduie 1: 77-164. 2004. Los procesos culturales del comienzo del Holo-ceno en la cuenca del Ebro y su contextualizacion. Sal-duie 4:17-40. CHANDLER H., SYKES B., ZILHÄO J. 2005. Using ancient DNA to examine genetic continuity at the Mesolithic-Neo-lithic transition in Portugal. In P. Arias, R. Ontanon, C. Garcia-Monco (eds.), Actas del III Congreso sobre el Neolitico en la Peninsula Ibérica. Instituto Internacional de Investigaciones Prehistoricas de Cantabria, Santander: 781-786. CORCHON M. S. 2006. Las cuevas de la Griega y Palome-ra (Ojo Guarena) y la cuestion de la cronologia del arte paleolitico en la Meseta. In F. Diez and G. Delibes (eds.), El Paleolitico Superior en la Meseta Norte espanola. Fun-dacion Studia Archaeologica 94: 75-111. CORTÉS M. (ed.) 2007. Cueva Bajondillo (Torremolinos). Secuencia cronocultural y paleoambiental del Cuater-nario reciente en la Bahia de Mälaga. Centro de Edicio-nes de la Diputacion de Malaga, Junta de Andalucia, Uni-versidad de Malaga, Fundacion Cueva de Nerja y Fundacion Obra Social de Unicaja. Malaga. CRUZ BERROCAL M., VICENT J. 2007. Rock art as an archaeological and social indicator: the Neolithisation of the Iberian Peninsula. Journal of Anthropological Archaeology 26: 676-697. DAVIS B. S. A, STEVENSON A. C. 2007. The 8.2 ka event and Early-Mid Holocene forest, fires and flooding in the Central Ebro Desert, NE Spain. Quaternary Science Reviews 26:1695-1712. DETRY C. 2007. Paleoecologia e paleoeconomia do Baixo Vale do Tejo no Mesolitico Final: o contributo dos mami-feros dos concheiros de Muge. Dissertaçao de Doutora-mento. Lisboa: Universidade Autonoma de Lisboa e Uni-versidade de Salamanca. DIEZ F., DELIBES G. 2006. El Paleoh'tico Superior en la Meseta Norte espanola. Fundacion Studia Archaeologica 94. ESQUEMBRE M. A., BORONAT J. D., JOVER F. J., MOLINA F. J., LUJAN A., FERNANDEZ J., MARTINEZ R., IBORRA P., FERRER C., RUIZ R., PÉREZ G., ORTEGA J. R. 2008. El yaci-miento neolitico del Barranquet de Oliva (Valencia). In M. Hernandez, J. Soler and J. A. Lopez (eds.), Actas del IV Congreso del Neolüico Peninsular (vol I). MARQ. Diputa-cion de Alicante, Alicante: 183-190. ESTREMERA M. S. 2003. Primeros agricultores y ganade-ros en la Meseta Norte. El Neolitico de la Cueva de La Vaquera (Torreiglesias, Segovia). Memorias d Arqueologia en Castilla y Leon, 11. Junta de Castilla y Leon. Zamora. FANO M. A. 2004. Un nuevo tiempo: El Mesolitico en la region Cantabrica. Kobie 8:337-402. FERNANDEZ-DOMINGUEZ E. 2005. Polimorfismos de DNA mitochondrial en poblaciones antiguas de la Cuenca Mediterrânea. PhD Dissertation. Universitat de Barcelona. FERNANDEZ LOPEZ DE PABLO J. 2006. Contribucion al conocimiento de la secuencia arqueologica y el habitat del Holoceno inicial en el Maestrazgo. Saguntum-PLAV 38: 23-47. FERNANDEZ LOPEZ DE PABLO J., GUILLEM P. M., MARTINEZ R., PÉREZ R. 2005. Nuevos datos sobre el Abric del Mas de Marti (Albocàcer, Castello). In P. Arias, R. Ontanon, C. Garcia-Monco (eds.), Actas del III Congreso sobre el NeoUtico en la Penmsula Ibérica. Instituto Internacional de Investigaciones Prehistoricas de Cantabria, Santander: 879-887. FERNANDEZ LOPEZ DE PABLO J., GOMEZ PUCHE M., DIEZ CASTILLO A., FERRER GARCIA C., MARTINEZ-ORTI A. 2008. Resultados preliminares del proyecto de investigacion sobre los origenes del Neolitico en el Alto Vinalopo y su co-marca: la revision de El Arenal de la Virgen (Villena, Alicante). In M. Hernandez, J. Soler and J. A. Lopez (eds.), Actas del IV Congreso del Neolüico Peninsular (vol I). MARQ. Diputacion de Alicante, Alacant: 107-116. FERNANDEZ LOPEZ DE PABLO J., GIBAJA J. F., PALOMO A. 2009. Geometric weapon elements during the Neolithic in the Eastern Iberian Peninsula: Typological, technological and functional aspects. Paleoethnologie 2008.1:328-340. FERNANDEZ LOPEZ DE PABLO J., JOCHIM M. A. in press. The impact of the 8200 cal yr BP climatic event on human mobility strategies during the Late Mesolithic in Iberia. Journal of Anthropological Research. FIEDEL S. T., ANTHONY D. W. 2007. Deerslayers, pathfinders and icemen. Origins of the European Neolithic as seen from the frontier. In M. Rockman and J. Steele (eds.), Colonization of Unfamiliar Landscapes: The archaeology of adaptation. Routledge. London and New York: 144163. FORTEA PÉREZ F. J. 1973. Los Complejos Microlaminares y Geométricos del Epipaleolüico Mediterrâneo Espa-nol. Universidad de Salamanca. FRIGOLA J., MORENO A., CACHO I., CANALS M., SIERRO F. J., FLORES J. A., GRIMALT J. O., HODELL D. A., CURTIS J. H. 2007. Holocene climate variability in the western Mediterranean region from a deepwater sediment record. Pale-oceanography 22: PA2209, doi:10.1029/2006PA001307. FUERTES N., NEIRA A. 2006. La cadena operativa primas-tica en el yacimiento mesolitico de "El Espertin" (Buron, Leon, Espana). In Roläo y Ferreira (eds.), Do Epipaleolf tico ao Calcolüico na Penmsula Ibérica. Actas do IV Con-gresso de Arqueologia Peninsular. Promontoria Monogra-fica, 4. Universidade do Algarve. Faro: 95-111. FUMANAL P. 1995. Los depositos cuaternarios en cuevas y abrigos. Implicaciones sedimentoclimaticas. In V. M. Ro-sello (ed.), El Cuaternario del Pa^s Valenciano. Universitat de València, València: 115-124. GAMBA C., FERNANDEZ E., OLIVER A., TIRADO M., BAEZA C., LOPEZ-PARRA A. M., ARROYO PARDO E. 2008. Population genetics and DNA preservation in ancient human remains from Eastern Spain. Forensic Science International: Genetics Supplement Series 1: 462-464. GARCIA-ARGÜELLES P., FULLOLA J. M. 2006. La Cueva del Parco (Alös de Balaguer, Lleida) y el Abrigo del Filador (Margalef de Montsant, Tarragona: Dos secuencia clave para el conocimiento del Epipaleolitico en el Nordeste Peninsular. In A. Alday (ed.), El mesolüico de muescasy denticulados en la Cuenca del Ebro y el litoral mediter-râneo peninsular. Diputacion Foral de Alava, Departa-mento de Cultura. Alava: 121-133. GARCIA GUIXÉ E., RICHARDS M. P., SUBIRÀ M. E. 2006. Palaeodiets of Humans and Fauna at the Spanish Mesoli-thic Site of El Collado. Current Anthropology 47 (3): 549556. GARCIA O., AURA J. E. 2006. El abric de la Falguera (Alcoi, Alacant): 8.000 anos de ocupacion humana en la cabecera del rio de Alcoi. Diputacion Provincial de Alicante. Alicante. GIRALT S., JULIÀ R. 2003. Water level reconstruction in closed lakes base on the mineral composition of sediments. In B. Valero-Garcés (ed.), Limnogeology in Spain: a Tribute to Kerry R. Kelts. Madrid: 305-325. GOMEZ M., FERNANDEZ LOPEZ DE PABLO J. 2009. Crono-logia absoluta por Carbono 14 del Epipaleolitico y Neoli- tico Antiguo en Villena. Memoria de las IV Ayudas a la Investigation. Fundacion Municipal José Maria Soler. Villena. Inédita. GONZÀLEZ-SAMPÉRIZ P., UTRILLA P., MAZO C., VALERO-GARCÉS B., SOPENA M. C., MORELLON M., SEBASTIÂN M., MORENO A., MARTINEZ M. 2009. Patterns of human occupation during the Early Holocene in the Central Ebro Basin (NE Spain) in response to the 8200-yr climatic event. Quaternary Research 71:121-132. HAAK W., FORSTER P., BRAMANTI B., MATSUMURA S., BRANDT G., TANZER M., VILLEMS R., RENFREW C., GRO-NENBORN D., ALT K. W., BURGER J. 2005. Ancient DNA from the First European Farmers in 7500-Years-Old Neolithic sites. Science 310:1016-1028. IRIARTE M. J. 2009. Vegetation landscape and the anthro-pization of the environment in the central sector of the Northern Iberian Peninsula: Current status. Quaternary International 200: 66-76. JACKES M., MEIKLEJOHN C. 2004. Building a method for the study of the Mesolithic-Neolithic transition in Portugal. In M. Budja (ed.), 11th Neolithic Studies. Documenta Praehistorica 31: 89-111. JIMENEZ-ESPEJO F. J., MARTINEZ F., SAKAMOTO T., IIJIMA K., GALLEGO D., HARADA N. 2007. Paleoenvironmental changes in the western Mediterranean since the last glacial maximum: High resolution multiproxy record from the Algero-Balearic basin. Palaeogeography, Palaeoclima-tology, Palaeoecology 246:292-306. JIMÉNEZ-GUIJARRO J. 1999. El proceso de neolitizacion del interior peninsular. In J. Bernabeu and T. Orozco (eds.), Actes del II Congrès del Neolitic a la Peninsula Ibé-rica. Saguntum-PLAVExtra 2: 493-502. 2005. El proceso de neolitizacion en la Comunidad de Madrid. In P. Arias, R. Ontanon, C. Garcfa-Moncö (eds.), Actas del III Congreso sobre el Neolitico en la Peninsula Ibérica. Instituto International de Investigaciones Prehistöricas de Cantabria, Santander: 903-912. 2008. El horizonte cardial del interior de la Peninsula Ibérica. In M. Hernandez, J. Soler and J. A. Lopez (eds.), Actas del IV Congreso del Neolitico Peninsular (vol II). MARQ. Diputaciön de Alicante, Alacant: 348-355. JIMÉNEZ-GUIJARRO J., ROJAS J. M., GARRIDO G., PERERA J. 2008. El yacimiento del Neolitico Inicial de La Paleta (Nu-mancia de la Sagra, Toledo). In M. Hernandez, J. Soler and J. A. Lopez (eds.), Actas del IV Congreso del Neolitico Peninsular (vol. I). MARQ. Diputaciön de Alicante, Ala-cant: 126-136. JUAN-CABANILLES J. 1990. Substrat épipaléolithique et néolithisation en Espagne: Apport des industries lithi- ques a l'identification des traditions culturelles. In D. Ca-hen, M. Otte (eds.), Rubané et cardial. Actes du Colloque de Liége (1988). Service de Préhistoire de l'Université de Liége. ERAUL 39. Liége: 417-435. JUAN-CABANILLES J., MARTI B. 2002. Poblamiento y pro-cesos culturales en la Peninsula Ibérica del VII al V mile-nio a.C. (8000-5500 BP). Una cartografia de la neolitizacion. In E. Badal, J. Bernabeu and B. Marti (eds.), Neolithic landscapes of the Mediterranean. Saguntum-PLAV Extra-5: 45-87. 2007-2008. La fase C del Epipaleolitico reciente: lugar de encuentro o linea divisoria. Reflexiones en torno a la neolitizacion en la fachada mediterranea peninsular. Veleia 24-25: 611-628. LALUEZA FOX C. 1996. Physical anthropological aspects of the Mesolithic-Neolithic transition in the Iberian Peninsula. Current Anthropology 37: 689-95. LOPEZ SÂEZ J. A., LOPEZ GARCIA P., CORTÉS SÂNCHEZ M. 2007. Paleovegetacion del Cuaternario reciente: Estudio ar-queopalinologico. In M. Cortés Sanchez (ed.), Cueva Bajon-dillo (Torremolinos). Secuencia cronocultural ypaleo-ambiental del Cuaternario reciente en la Bahia de Mâla-ga. Centro de Ediciones de la Diputacion de Malaga, Junta de Andalucia, Universidad de Mälaga, Fundacion Cueva de Nerja y Fundacion Obra Social de Unicaja. Malaga: 139-156. LUBELL D., JACKES M., SCHWARCZ H., KNYF M., MEIKLEJOHN C. 1994. The Mesolithic-Neolithic transition in Portugal: isotopic and dental evidence of diet. Journal of Archaeological Science 21: 201-216. LUBELL D., JACKES M., SHEPPARD P., ROWLEY-CONWY P. 2007. The Mesolithic-Neolithic in the Alemtejo: archaeological investigations, 1984-1986. In IV Congresso de Ar-queologia Peninsular. From the Mediterranean basin to the Portuguese Atlantic shore: papers in honor of Anthony Marks. Faro: Universidade do Algarve. Promonto-ria Monografica 7: 209-230. MANEN C., MARCHAND G., CARVALHO A. F. 2007. Le Néolithique ancien de la Péninsule Ibérique: vers une nouvelles evaluation du mirage africain? In J. Évin (ed.), Un siècle de construction du discours scientifique en Préhistoire 3:133-151. MARTI B. 2008. Cuevas, poblados y santuarios neoliticos: Una perspectiva mediterranea. In M. Hernandez, J. Soler and J. A. Lopez (eds.), Actas del IV Congreso del Neolitico Peninsular (vol I). MARQ. Diputacion de Alicante, Alacan-te: 17-27. MARTI B., PASCUAL V., GALLART M. D., LOPEZ P., PÉREZ M., ACUNA J. D., ROBLES F. 1980. Cova de l'Or (Beniarrés, Alicante), Vol. II. Trabajos Varios del S.I.P., 65. Dipu-tacion de Valencia. Valencia. MARTINEZ-CORTIZAS A., COSTA-CASAIS M. and LOPEZ-SAEZ J. A. 2009. Environmental change in NW Iberia between 7000 and 5000 cal BC. Quaternary International 200: 77-89. MARTINEZ-MORENO J., MORA R., CASANOVA J. 2006. El Mesolitico de los Pirineos surorientales: una reflexion sobre el significado de las "facies de fortuna" del Postgla-ciar. In A. Alday (ed.), El Mesolitico de muescas-denticu-lados en la Cuenca del Ebro y el Litoral Mediterrâneo peninsular. Memorias de Yacimientos Alaveses, 11, Diputacion Foral de Älava: 158-188. MARTINS J. M. M., CARVALHO A. F., SOARES A. M. M. 2008. A calibraçao das datas de radiocarbono dos esqueletos hu-manos de Muge. Promontoria 6: 73-93. MCCLURE S. B., BARTON C. M., JOCHIM M. A. 2009. Human behavioral ecology and climate change during the transition to agriculture in Valencia, Eastern Spain. Journal of Anthropological Research 65(2): 253-269. MCKEEVER M.1984. Comparative palynological studies of two lake sites in Western Ireland and Northwestern Spain. Msc Thesis, Trinity College, Dublin. MESTRES J. 1992. Neolitizacio i territori. In M. Cura Mo-rera (ed.), Estat de la investigaciô sobre elNeolitic a Ca-talunya. 9è Colloqui Internacional d'Arqueologia de Puigcerdà: 72-75. 1995. La datacio per radiocarboni i el calibratge de les dates radiocarböniques. Objectiu, problemes i aplica-cions. Revista d'Arqueologia de Ponent 5:260-275. MIRET C., MORALES J. V., PEREZ M., GARCIA O., AURA J. E. 2007. Els materials mesolitics de la Cova del Mas del Gelat (Alcoi, l' Alcoià, Alacant). Recerques del Museu d'Alcoi 15: 7-18. MONTEIRO-RODRIGUES S. 2003. Comment on Carvalho -A emergência do Neolitico no actual territorio português: pressupostos teoricos, modelos interpretativos e a evidên-cia empirica. O Arqueôlogo Português 21:118-125. MONTES L., UTRILLA P., MAZO C. 2006. El Epipaleohtico Macrolitico en Aragon en el contexto del Valle del Ebro y la Cataluna costera. In A. Alday (ed.), El Mesolitico de muescas y denticulados en la Cuenca del Ebro y el lito-ral mediterrâneo peninsular. Diputacion Foral de Älava, Departamento de Cultura. Alava: 193-219. MORELLON M., VALERO B., MORENO A., GONZÄLEZ P., MATA P., ROMERO O., MAESTRO M., NAVAS A. 2007 Holocene palaeohydrology and climate variability in northeastern Spain: The sedimentary record of Lake Estanya (Pre-Pyrenean range). Quaternary International 181: 1531. MUNOZ C., RAMIL P., RODRIGUEZ M. 2001. Datos sobre la dinamica del paisaje vegetal en el Parque Natural del Lago de Sanabria (Zamora, NO de la PI) durante el final del Pleistoceno. In M. A. Fombella, D. Fernandez and R. M. Valencia (eds.), Palinologia: Diversidad y aplicacio-nes. Universidad de Leon. Leon: 71-78. OLÀRIA C., GUSI F., GOMEZ J. L. 2005. Un enterramiento Meso-Neolitico en el Cingle del Mas Nou (Ares del Maestre, Castellon) del 7000 BP en territorio de arte levantino. In P. Arias, R. Ontanon, C. Garcia-Monco (eds.), Actas del III Congreso sobre el Neolitico en la Peninsula Ibérica. In-stituto Internacional de Investigaciones Prehistoricas de Cantabria, Santander: 615-623. ORTEGA A. I., JUEZ L., CARRETERO J. M., ORTEGA M. C., ARSUAGA J. L., PEREZ A. 2008. El Neolitico en la nueva secuencia estratigrafica del yacimiento del Portalon de Cueva Mayor (Sierra de Atapuerca, Burgos). In M. Hernandez, J. Soler, J. A. Lopez (eds.), Actas del IV Congreso del Neolitico Peninsular (vol. 1). MARQ. Diputacion de Alicante, Alicante: 221-229. PENA-CHOCARRO L., ZAPATA L., IRIARTE M. J., GONZÄLEZ M., STRAUS L. G. 2005. The oldest agriculture in northern Atlantic Spain: new evidence from El Miron Cave (Ramales de la Victoria, Cantabria). Journal of Archaeological Science 32:579-587. PICAZO J. V., RODANÉS J. M. 2008. Talleres de silex, pobla-dos y aldeas. Una cabana mesohtica en el Cabezo de la Cruz (La Muela, Zaragoza). In M. Hernandez, J. Soler, J. A. Lopez (eds.), Actas del IV Congreso del Neolitico Peninsular (vol. 1). MARQ. Diputacion de Alicante, Alicante: 137-142. PINHASI R., PLUCIENNIK M. 2004. A Regional Biological Approach to the Spread of Farming in Europe. Current Anthropology 45:59-82. RAMOS J., LAZARICH M. 2002. El asentamiento de 'El Re-tamar' (Puerto Real, Cadiz). Contribution al estudio de la Formation Social Tribal y a los inicios de la economia de production en la Bahia de Cadiz. Servicio de Publicacio-nes de la Universidad de Cadiz. Cadiz. RAMOS J., CASTANEDA V. (eds.) 2005. Excavaciôn en el asentamiento prehistôrico del Embarcadero del rio Pal-mones (Algeciras, Câdiz): una nueva contribuciôn al estudio de las ultimas comunidades cazadoras y reco-lectoras. Universidad de Cadiz, Servicio de Publicaciones. Fundacion Municipal de Cultura 'José Luis Cano'. Algeciras. RAMOS J., LAZARICH M., CÄCERES I., PÉREZ M., CASTANEDA V., HERERO N., DOMINGUE Z. S. 2005. El asentamiento de El Retamar: sintesis del registro arqueologico y enmarque socioeconomico e historico". In P. Arias, R. Ontanon, C. Garcia-Monco (eds.), Actas del III Congreso sobre el Neolitico en la Peninsula Ibérica. Instituto International de Investigaciones Prehistoricas de Cantabria, Santander: 509-518. REIMER P. J., BAILLIE M. G. L., BARD E., BAYLISS A., BECK J. W., BERTRAND C. J. H. 2004. IntCal04 terrestrial radiocarbon age calibration, 0-26 cal kyr BP. Radiocarbon 46 (3): 1029-1058. ROCHE J. 1972. Le gisement mésolithique de Moita do Sebatiâo Muge Portugal. I Archéologie. Instituto de Alta Cultura, Lisbon. ROJO M., KUNST M. 1999. El Valle de Ambrona: un ejem-plo de la primera colonization Neolitica de las tierras del Interior Peninsular. In J. Bernabeu and T. Orozco (eds.), Actes del II Congrès del Neolitic a la Peninsula Ibérica. Saguntum-PLAV Extra 2: 259-270. ROJO M., KUNST M., GARRIDO R., GARCIA I. 2006. La neo-litizacion de la Meseta Norte a la luz del C-14: Analisis de 47 dataciones absolutas inéditas de dos yacimientos domésticos del Valle de Ambrona, Soria, Espana. Archivo de Prehistoria Levantina XXVI: 39-100. ROJO M., GARRIDO R., GARCIA I. 2008. La ocupacion del Neolitico antiguo del Abrigo de Carlos Alvarez/La Dehesa (Mino de Medinaceli, Soria). In M. Hernandez, J. Soler, J. A. Lopez (eds.), Actas del IV Congreso del Neolitico Peninsular (vol. 1). MARQ. Diputacion de Alicante, Alicante: 246-251. ROKSANDIC M. 2006. Analysis of burials from the new excavations of the sites Cabeço da Amoreira and Cabeço da Arruda (Muge, Portugal). In Roläo y Ferreira (eds.), Do Epipaleolitico ao Calcolitico na Peninsula Ibérica. Actas do IV Congresso de Arqueologia Peninsular. Promonto-ria Monografica 4. Universidade do Algarve. Faro: 43-54. RUIZ J., SMITH P. 2001. El yacimiento de Cubio Redondo (Matienzo, Ruesga). Munibe 53:31-55. 2003. La Cueva de Cofresnedo en el valle de Matienzo. Actuaciones arqueolôgicas 1996-2001. Gobierno de Cantabria, Santander. SAMPIETRO M. L., LAO O., CARAMELLI D., LARI M., POU R., MARTI M., BERTRANPETIT J., LALUEZA-FOX C. 2007. Pa-laeogenetic evidence supports a dual model of Neolithic spreading into Europe. Proceedings of the Royal Society B 274: 2161-2167. SHENNAN S. J. 2007. The spread of farming into Central Europe and its consequences: evolutionary models. In T. Kohler and S. E. van der Leeuw (eds.), The Model-Based Archaeology of Socionatural Systems. SAR Press, Santa Fe: 141-156. SOARES A. M. M. 1993. The 14C content of marine shells: evidence for variability in coastal upwelling off Portugal during the Holocene. Isotope Techniques in the Study of Past and Current Environmental Changes in the Hydrosphere and the Atmosphere (Proceedings). IAEA. Vienna: 471-485. SOARES A. M. M., DIAS J. M. A. 2006. Coastal upwelling and radiocarbon-evidence for temporal fluctuations in ocean reservoir effect off Portugal during the Holocene. Radiocarbon 48 (1): 45-60. SOLER J., FERRER C., ROCA DE TOGORES C., GARCIA G. 2008. Cova d'En Pardo (Planes, Alicante). In M. Hernandez, J. Soler, J. A. Lopez (eds.), Actas del IVCongreso del Neolitico Peninsular (vol. 1). MARQ. Diputacion de Alicante, Alicante: 79-89. STINER M. C., BICHO N. F., LINDLY J., FERRING C. R. 2003. Mesolithic to Neolithic transitions: new results from shell-middens in the western Algarve, Portugal. Antiquity 77 (295): 75-86. STIKA H. P. 2005. Early Neolithic agriculture in Ambrona, Provincia Soria, central Spain. Vegetation History and Archaeobotany 14:189-197. STRAUS L. G. 2004. Transitions: into and out of Mesolithic adaptations along the Atlantic façade of Europe and beyond. In G. A. Clark and M. Gonzalez-Morales (eds.), The Mesolithic of the Atlantic Façade: proceedings of the Santander Symposium. Arizona State University. Anthropological Research Papers 55: 249-260. 2008. The Mesolithic of Atlantic Iberia. In G. Bailey and P. Spikins (eds.), Mesolithic Europe. Cambridge University Pres, Cambridge: 302-327. TAPIA J., ALVAREZ E., CUBAS M., CUETO M., ETXEBERRIA F., GUTIERREZ I., HERRASTI L., RUIZ M. 2008. La Cueva de Linatzeta (Lastur, Deba, Gipuzkoa). Un nuevo contexto para el estudio del Mesolitico en Gipuzkoa. Munibe 59: 119-131. UMBELINO C. 2006. Outros sabores do Passado. As anâ-lises de oligoelementos e de isôtopos estâveis na recon-struçâo da dieta das comunidades humanas do Mesolitico Final e do Neolitico Final/Calcolitico do territôrio português. Dissertaçao de Doutoramento. Coimbra: Facul-dade de Ciências e Tecnologia, Universidade de Coimbra. UTRILLA P. 2005. Arte rupestre en Aragon. 100 anos des-pués de Calapata. In M. Hernandez and J. Soler (eds.), Arte rupestre en la Espana mediterrânea. Alicante: 341378. UTRILLA P., CAVA A., ALDAY A., BALDELLOU V., BARAN-DIARAN I., MAZO C., MONTES L. 1998. Le passage du Meso- lithique au Néolitique ancien dans le bassin de l'Ebre (Espagne) d'après les datacions C14. Préhistoire Européenne 12:171-194. UTRILLA P., RODANÉS J. M. 2004. Un asentamiento Epi-paleolûico en el valle del rto Martm. El abrigo de los Earns (Arino, Teruel). Monografias Arqueologicas 39. Uni-versidad de Zaragoza. Zaragoza. UTRILLA P., LORENZO J. I., BALDELLOU V., SOPENA M. C., AYUSO P. 2008. Enterramiento masculino en fosa, cubier-to de cantos rodados, en el neolitico antiguo de la Cueva de Chaves. In M. Hernandez, J. Soler, J. A. Lopez (eds.), Actas del IV Congreso del Neolüico Peninsular (vol. II). MARQ. Diputacion de Alicante, Alacant: 131-140. VAN DER SHRIECK T., PASSMORE D. G., ROLÄO J., STEVENSON A. C. 2007. Estuarine-fluvial floodplain formation in the Holocene Lower Tagus valley (central Portugal) and implications for Quaternary fluvial system evolution. Quaternary Science Reviews 26:2937-2957. VAN DER SHRIECK T., PASSMORE D. G., FRANCO F., STEVENSON A. C., BOOMER I., ROLÄO J. 2008. Holocene pa-laeoecology and floodplain evolution of the Muge tributary, Lower Tagus Basin, Portugal. Quaternary International 189:135-151. VAQUERO 2006. El Mesolitico de facies macrolitica en el centro y sur de Cataluna. In A. Alday (ed.), El Mesolüico de muescas y denticulados en la Cuenca del Ebro y el litoral mediterrâneo peninsular. Diputacion Foral de Älava, Departamento de Cultura: 137- 160. VAZQUEZ J. M. 2004. The Mesolithic on the northwest coast of the Iberian Peninsula. In G. A. Clark and M. Gonzalez-Morales (eds.), The Mesolithic of the Atlantic Façade: proceedings of the Santander Symposium. Arizona State University. Anthropological Research Papers 55: 103-110. VERGÈS J. M., ALLUÉ E., ANGELUCCI D. E., BURJACHS F., CARRANCHO A., CEBRIÀ A., EXPOSITO I., FONTANALS M., MORAL S., RODRIGUEZ A., VAQUERO M. 2008. Los niveles neoliticos de la Cueva del El Mirador (Sierra de Ata-puerca, Burgos): Nuevos datos sobre la implantacion y el desarrollo de la economia agropecuaria en la Submeseta Norte. In M. Hernandez, J. Soler and J. A. Lopez (eds.), Actas del IV Congreso del Neolüico Peninsular (vol. I). MARQ. Diputacion de Alicante, Alacant: 418-427. VICENT GARCIA J. M. 1997. The Island Filter Model Revisited. In M. S. Balmuth, A. Gilman, L. Prados Torreira (eds.), Encounters and Transformations: The Archaeology of Iberia in Transition. Sheffield Academic Press, Sheffield: 1-13. VICENTE M., MARTINEZ R., GUILLEM P. M., IBORRA M. P. 2009. El Cingle del Mas Cremat (Portell de Morella, Ca- stello). Un asentamiento en altura con ocupaciones del Mesolitico Reciente. In II Reunion sobre Mesolüico de la Cuenca del Ebro y Litoral Mediterrâneo. Universidad de Zaragoza. Zaragoza: 273-286. WELLS S. 2007. Deep ancestry. Inside the genographic project. National Geographic, Washington DC. WENINGER B., ALRAM-STERN E., BAUER E., CLARE L., DANZEGLOCKE U., JÖRIS O., KUBATZKI C., ROLLEFSON G., TODOROVA H., VAN ANDEL T. 2006. Climate forcing due to the 8200 cal yr BP event observed at Early Neolithic sites in the Eastern Mediterranean. Quaternary Research 66 (3): 401-420. ZAPATA L. 1995. El yacimiento arqueologico de la Cueva de Pico Ramos (Muskiz, Bizkaia). In Primeros agriculto-res y ganaderos en el Cantâbrico y Alto Ebro. Cuader-nos de Prehistoria - Arqueolog^a 6:251-257. 2002. Origen de la agricultura en el Pa^s Vasco y transformaciones en elpaisaje. Kobie Anejo 4, Bilbao. ZAPATA L., PENA-CHOCARRO L., PEREZ-JORDA G., STIKA H. P. 2004. Early Neolithic Agriculture in the Iberian Peninsula. Journal of World Prehistory 18 (4): 283-325. ZAPATA L., BALDELLOU V., UTRILLA P. 2008. Bellotas de cronologia neolitica para consumo humano en la Cueva de Chaves (Bastaras, Huesca). In M. Hernandez, J. Soler and J. A. Lopez (eds.), Actas del IV Congreso del Neolüico Peninsular (vol. I). MARQ. Diputacion de Alicante, Ala-cant: 402-410. ZILHÄO J. 1992. Gruta do Caldeiräo. O Neolüico antigo. Trabalhos de Arqueologta, 6. Instituto Portugues do Patri-monio Arquitectonico e Arqueologico. Lisboa. 1993. The spread of agro-pastoral economies across mediterranean Europe: a view from the far west. Journal of Mediterranean Archaeology 6 (1): 5-63. 1997. Maritime pioneer colonization in the Early Neolithic of the West Mediterranean. Testing the model against the evidence. Poročilo o raziskovanju paleolitika, neolüika in eneolitika v Sloveniji XXIV: 19-42. 2000. From the Mesolithic to the Neolithic in the Iberian Peninsula. In T. D. Price (ed.), Europe's firstfarmers. Cambridge University Press. Cambridge: 144182. 2001. Radiocarbon evidence for maritime pioneer colonization at the origins of farming in west Mediterranean Europe. Proceedings of the National Academy of Sciences 98:14180-14185. 2003. Algumas observaçôes acerca do Mesolitico do interior peninsular e do modelo de passagem ao Neoli- tico através de colonizaçao pioneira por via maritima. Comments on Carvalho - A emergência do Neolitico no actual territorio português: pressupostos teoricos, modelos interpretativos e a evidência empirica. O Arqueologo Português 21:108-118. 2004. Comments on Pinhasi and Pluciennick - A Regional Biological Approach to the Spread of Farming in Europe. Current Anthropology 45: 77-78. ZVELEBIL M. 2000. The Social Context of the Agricultural Transition in Europe. In C. Renfrew and K. Boyle (eds.), Archaeogenetics: DNA and the population prehistory of Europe. McDonald Institute for Archaeological Research Monographs. Cambridge: 57-79.