Rendering Death: Ideological and Archaeological Narratives from Recent Prehistory (Iberia): Proceedings of the conference held in Abrantes, Portugal, 11 May 2013 9781407312873, 9781407342535

This book offers a perspective on death and memory in recent Prehistory on the western Iberian Peninsula (Portugal, Span

160 78 24MB

English Pages [145] Year 2014

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Rendering Death: Ideological and Archaeological Narratives from Recent Prehistory (Iberia): Proceedings of the conference held in Abrantes, Portugal, 11 May 2013
 9781407312873, 9781407342535

Table of contents :
Front Cover
Title Page
Copyright
Table of Contents
List of Contributors
Rendering Death - Ideological and archaeological speeches from recent Prehistory (Iberia)
Custodian stones: human images in the megalithism of the Southern Iberian Peninsula
The contribution of Manuel Heleno to the knowledge of the funerary Megalithic in Alentejo
Megalithic Rites of North Alentejo – Portugal
Death as “life’s mirror”? Funerary practices and trajectories of complexity in the prehistory of peasant societies of Iberia
The mound at Cimo dos Valeiros (Serra Vermelha, Oleiros, Castelo Branco). A Neolithic burial site in the Central Cordillera, south of Serra da Estrela
Caves, Megalithism and Tumuli – Three diachronic realities in funerary archaeography from Alto Ribatejo –
Collective Burial Caves in Spanish Extremadura: Chronology, Landscapes and Identities
Between dead and alive - the recent prehistory of the municipality of Pampilhosa da Serra (Central Portugal)
Between norm and variation in the semiotic of the funerary world: examples and discussion of some abnormal graves in the Bronze Age Europe
The polimorphism of graves and the distribution of archeological remains in the Southwest Bronze Age necropolis of Soalheironas (Alcoutim)
The faces of death: from the Bronze to the Iron Age, between the North and the South of the Portuguese territory

Citation preview

Rendering Death: Ideological and Archaeological Narratives from Recent Prehistory (Iberia) Proceedings of the conference held in Abrantes, Portugal, 11 May 2013

Edited by

Ana Cruz Enrique Cerrillo-Cuenca Primitiva Bueno Ramírez João Carlos Caninas Carlos Batata

BAR International Series 2648 2014

Published in 2016 by BAR Publishing, Oxford BAR International Series 2648 Rendering Death: Ideological and Archaeological Narratives from Recent Prehistory (Iberia) © The editors and contributors severally and the Publisher 2014 The authors' moral rights under the 1988 UK Copyright, Designs and Patents Act are hereby expressly asserted. All rights reserved. No part of this work may be copied, reproduced, stored, sold, distributed, scanned, saved in any form of digital format or transmitted in any form digitally, without the written permission of the Publisher.

ISBN 9781407312873 paperback ISBN 9781407342535 e-format DOI https://doi.org/10.30861/9781407312873 A catalogue record for this book is available from the British Library BAR Publishing is the trading name of British Archaeological Reports (Oxford) Ltd. British Archaeological Reports was first incorporated in 1974 to publish the BAR Series, International and British. In 1992 Hadrian Books Ltd became part of the BAR group. This volume was originally published by Archaeopress in conjunction with British Archaeological Reports (Oxford) Ltd / Hadrian Books Ltd, the Series principal publisher, in 2014. This present volume is published by BAR Publishing, 2016.

BAR

PUBLISHING BAR titles are available from:

E MAIL P HONE F AX

BAR Publishing 122 Banbury Rd, Oxford, OX2 7BP, UK [email protected] +44 (0)1865 310431 +44 (0)1865 316916 www.barpublishing.com

Contents LIST OF CONTRIBUTORS .............................................................................................................................................................. II LIST OF FIGURES, MAPS, TABLES AND GRAPHICS.............................................................................................................. IV RENDERING DEATH - IDEOLOGICAL AND ARCHAEOLOGICAL SPEECHES FROM RECENT PREHISTORY (IBERIA)

ANA CRUZ ......................................................................................................................................................................... 1  CUSTODIAN STONES: HUMAN IMAGES IN THE MEGALITHISM OF THE SOUTHERN IBERIAN PENINSULA.

P. BUENO RAMIREZ, R. DE BALBÍN BEHRMANN, R. BARROSO BERMEJO ........................................................................................ 3  THE CONTRIBUTION OF MANUEL HELENO TO THE KNOWLEDGE OF THE FUNERARY MEGALITHIC IN ALENTEJO

LEONOR ROCHA ................................................................................................................................................................ 13  MEGALITHIC RITES OF NORTH ALENTEJO – PORTUGAL

JORGE DE OLIVEIRA ............................................................................................................................................................ 23  DEATH AS “LIFE’S MIRROR”? FUNERARY PRACTICES AND TRAJECTORIES OF COMPLEXITY IN THE PREHISTORY OF PEASANT SOCIETIES OF IBERIA.

JOÃO CARLOS SENNA‐MARTINEZ ..................................................................................................................................... 35  THE MOUND AT CIMO DOS VALEIROS (SERRA VERMELHA, OLEIROS, CASTELO BRANCO). A NEOLITHIC BURIAL SITE IN THE CENTRAL CORDILLERA, SOUTH OF SERRA DA ESTRELA

JOÃO CARLOS CANINAS, MÁRIO MONTEIRO, ANDRÉ PEREIRA, EMANUEL CARVALHO, FRANCISCO HENRIQUES,   JOÃO ARAÚJO GOMES & LÍDIA FERNANDES, ÁLVARO BATISTA ................................................................................................... 45  CAVES, MEGALITHISM AND TUMULI – THREE DIACHRONIC REALITIES IN FUNERARY ARCHAEOGRAPHY FROM ALTO RIBATEJO –

ANA CRUZ, ANA GRAÇA, LUIZ OOSTERBEEK ........................................................................................................................... 61  COLLECTIVE BURIAL CAVES IN SPANISH EXTREMADURA: CHRONOLOGY, LANDSCAPES AND IDENTITIES

ENRIQUE CERRILLO CUENCA, ANTONIO GONZÁLEZ CORDERO .................................................................................................... 77  BETWEEN DEAD AND ALIVE - THE RECENT PREHISTORY OF THE MUNICIPALITY OF PAMPILHOSA DA SERRA (PORTUGAL CENTER)

CARLOS BATATA, FILOMENA GASPAR .................................................................................................................................... 91  BETWEEN NORM AND VARIATION IN THE SEMIOTIC OF THE FUNERARY WORLD: EXAMPLES AND DISCUSSION OF SOME ABNORMAL GRAVES IN THE BRONZE AGE EUROPE

DAVIDE DELFINO ............................................................................................................................................................. 105  THE POLIMORPHISM OF GRAVES AND THE DISTRIBUTION OF ARCHEOLOGICAL REMAINS IN THE SOUTHWEST BRONZE AGE NECROPOLIS OF SOALHEIRONAS (ALCOUTIM)

JOÃO LUÍS CARDOSO, ALEXANDRA GRADIM ......................................................................................................................... 119  THE FACES OF DEATH: FROM BRONZE TO IRON AGE, BETWEEN THE NORTH AND THE SOUTH OF THE PORTUGUESE TERRITORY

RAQUEL VILAÇA .............................................................................................................................................................. 125 



 

List of Contributors Alexandra Gradim

Archaeologist of the Alcoutim Municipal Council

Álvaro Batista

Archaeologist Member or collaborator with Upper Tagus Study Association (AEAT)

Ana Cruz

Prehistory Centre from Polytechnic Institute of Tomar (Portugal) “Quaternary and Prehistory” group of the Geosciences Centre from Coimbra University (uID73- F.C.T.). [email protected]

Ana Graça

Prehistory Centre from Polytechnic Institute of Tomar (Portugal) “Quaternary and Prehistory” group of the Geosciences Centre from Coimbra University (uID73- F.C.T.). [email protected]

André Pereira

Archaeologist Member or collaborator with Upper Tagus Study Association (AEAT)

Antonio González Cordero

Archaeologist Professor in IES Zurbarán de Navalmoral de la Mata (Cáceres) Cultural advisor at Fundación cultural Antonio Concha

Carlos Batata

CEO at OZECARUS, Serviços Arqueológicos, Lda Serra-Mãe, Cultural Association of Beira Serra

Davide Francesco Delfino

Earth and Memory Institute (ITM) Abrantes municipality – M.I.A.A project. “Quaternary and Prehistory” group of the Geosciences Centre from Coimbra University (uID73- F.C.T.)

Emanuel Carvalho

Archaeologist Member or collaborator with Upper Tagus Study Association (AEAT)

Enrique Cerrillo-Cuenca

“Ramón y Cajal” Fellow, Instituto de Arqueología – Mérida (CSIC – GobEx)

Filomena Gaspar

Technician at Abrantes municipality. Serra-Mãe, Cultural Association of Beira Serra

Francisco Henriques

Archaeologist Member or collaborator with Upper Tagus Study Association (AEAT)

João Araújo Gomes

Archaeologist Collaborator with Upper Tagus Study Association (AEAT) Centre of Geographical Studies, Institute of Geography and Territorial Planning, University of Lisbon

ii 

 

João Carlos Caninas

Archaeologist Master’s degree in Archaeology (Faculty of Letters, University of Oporto Doctoral candidate in Archaeology (CHAIA – University of Évora) Member of the Upper Tagus Study Association (AEAT) Managing director of EMERITA Lda

João Carlos Senna-Martinez

Centro de Arqueologia (Uniarq) da Universidade de Lisboa. 1600-214 LISBOA. PORTUGAL. [email protected]

João Luís Cardoso

Universidade Aberta and Centre of Archaeological Studies of Oeiras Council (Oeiras Municipal Council)

Jorge de Oliveira

CHAIA (University of Évora, Portugal).

Leonor Rocha

CHAIA Researcher School of Social Sciences - University of Évora

Lídia Fernandes

Archaeologist Member or collaborator with Upper Tagus Study Association (AEAT)

Luiz Oosterbeek

Polytechnic Institute of Tomar (Portugal) Earth and Memory Institute (ITM) “Quaternary and Prehistory” group of the Geosciences Centre from Coimbra University (uID73- F.C.T.)

Mário Monteiro

Archaeologist Member or collaborator with Upper Tagus Study Association (AEAT)

Primitiva Bueno Ramirez

Area de Prehistoria Alcalá University Spain [email protected]

Raquel Vilaça

Institute of Archaeology from Faculty of Letters Department of History, Archaeology and Arts, Coimbra University CEAUCP/FCT [email protected]

Rodrigo de Balbín Behrmann

Area de Prehistoria Alcalá University Spain [email protected]

Rosa María Barroso Bermejo

Area de Prehistoria Alcalá University Spain [email protected]

iii 

 

iv 

 

Rendering Death - Ideological and archaeological speeches from recent Prehistory (Iberia) Ana Cruz

ideological conceptions of death, maintaining the architectonic form of mound but invisible in the landscape.

Editorial Death causes in the human being a vulnerability, which added the fear of the unknown, induces a singular behavior towards the deceased relative.

The reuse of megalithic monuments is not unique also the caves are used during this period. The natural karstic cavities are interesting because of their relation with human sensorial activity. The access to these spaces carries a psychological and emotional adaptation to light differences and orientation once they are dark spaces and with different humidity and temperature comparing with outer spaces.

Parallel to this feelings comes a will to organize systematically a place for the dead, either ostensible, as in megalithic monuments, either latent in the landscape, as the caves-necropolis, cists and tumuli. Finite and the facts towards it in recent Prehistory

The study of archaeography of past contexts gives us one perception of how the system of beliefs and the symbolic universe might have been built creating a form of understanding about recent prehistory communities’ way of dealing with death.

In societies with diverse patterns death is seen in an individual perspective. Nevertheless, in Neolithic and Calcolithic it appears as collective contexts. Dealing with death must have had forms of religiosity in which the dead were increasingly relevant, as observed in the corps treatment. However, the data from archaeography show that the body was always important, even if buried collectively in a pit or ossuary.

We may induce an active role of the parents and other community members within the funerary events; since the moment of death until the final deposition of the body. Many other considerations can be made. These happened in the event that took place under the theme “Rendering Death - Ideological and archaeological speeches from recent Prehistory (Iberia)”

It seems, in Megalithism, during Neolithic and Calcolithic, that the common ancestor was being worshiped, through its reception in the chamber or corridors; whether when we face an isolate individual we revere the person. Possibly this last reality is due to a system where the person is the symbol of parenthood, so their heirs can claim proprieties.

References BRADLEY, R. (1993) – Altering the Earth. The origins of monuments in Britain and Continental Europe. Ediburgh: Society of Antiquaries of Scotland.

If we choose to consider funerary monuments as memory keepers (collective or individual) we may be towards a situation as referred to by Bradley (1993) with difference between utilitarian and ritual time. The ritual will be a non-measurable time, however, forces a line of continuity between past and present being the messenger of tradition and culture. In the climax of Megalithism there is a sense or message transmitted through monumental constructions, probably meaning a territorial claim by Neolithic agro-pastoral communities. If by one hand, was an act of impose ownership over the territory also their density and visibility have transformed and created a new landscape. The Bronze Age communities may have inherited a system of believes based in Neolithic simbolism as well as its territory management system. This link was strong enough to guarantee some acts such as burials in megalithic monuments. Is quite likely that the sense of ownership was reflected in oral tradition and consequently in the symbolic relation with death rituals. Thus, a new perspective towards death comes with the ritual of cist inhumation and incineration, with new 1 

 

 



 

Custodian stones: human images in the megalithism of the Southern Iberian Peninsula P. Bueno Ramirez, R. de Balbín Behrmann, R. Barroso Bermejo

If we consider the mobile art, menhirs, stelae and statues the Iberian Peninsula is one of the areas in the Atlantic megalithism which contains the most antropomorphic images associated to mortuary rituals.

Introduction In Atlantic Europe the construction of megaliths is a constant fact associated to Neolithic settling. The human remains that are kept in these structures are surrounded by stone materialisations of old ancestors. Sometimes the stones are cut into human shapes, but also antropomorphic engravings and paintings with a human shape occur. The stone and the human body are part of the same ideological discourse which justifies the identification of megalithic stones as antropomorphic representations. The tendency to emulate natural shapes can be explained by the search for a connection between nature and the products and human culture. This emulation is expressed by establishing identities among specific geologies and characters or megalithic structures and natural caves. Therefore the symbiosis of stone/human body is a plausible interpretation for the moving of big stones. This moving phenomenon is a part of the cultural expressions associated to the development of European societies from the VI to the III millennium cal BC (Bueno et. al., 2008).

In this context the recently acquired information from the South proposes important reflections regarding the connection between mobile pieces and the development of megalithic statuary. In fact, a great number of antropomorphic figurines clearly define identitary expressions (Bueno, 1992, 2010; Hurtado, 2008). Many of these expressions were made of highly valuable raw materials (Valera, 2010a). They are interesting because they were produced in the most abundant area regarding the density of information about schematic art. Therefore, megalithic art and schematic art converge in the same areas (Bueno & Balbín, 2006). Among these novelties two matters stand out: the continuous presence of menhirs throughout the whole megalithic diachrony and the remarkable ideological intensification which is associated to megalithic structures as of the IV millennium cal BC. This leads to several stelae in very visible positions, associated to different types of architectures (Bueno et. al., 2007a). Their iconographic relationships indicate a varied network of cultural interactions. Evidences of this network are also found in the prestige products, associated to these monuments’ gravegoods. Thus it seems to be coherent to interpret these images as identifying and protective elements of the sepulchres.

The identification of the stones used in megalithic structures as human images is preceded by a long tradition in European historiography. The best known interpretation is the one applied to the “goddesses” placed at the entrance of several French hypogeums. Arnal (1976) extended this hypothesis to the menhir‐statues placed at the entrance of other funerary structures in the South of France. L’Helgouach (1986 and 1993: 11) has suggested this interpretation for the reused stones with antropomorphic decorations in megaliths in the West of France. The amount of Breton examples and their spectacular nature are some of the reasons why this area is considered to be especially important during the development of the Atlantic megalithism (Cassen et. al., 2009; Cassen, 2007). Nonetheless, in the past years the Iberian Peninsula has contributed an unprecedented amount of cases which provide further interpretations regarding these stones, guarding death (Bradley, 2009; Bueno, 1995; Bueno et. al., 2005 & 2007a).

Their presence in the South of the Peninsula might be explained by the intensification of research during the last years. However, it is a fact that their relationship with an abundant panorama of mobile representations is an unavoidable reference for their analysis. Matters of terminology J.Guilaine already pointed out a long time ago that an overly divided terminology for the Atlantic megalithism was causing problems. The same happens when we try to analyse human representations (Bueno, 1995; Bueno et. al., 2003 & 2005). It is especially difficult to establish strong criteria in order to define and separate different versions. A big part of the problems comes from the historiographical tradition.

The human figures from Iberian megalithism present a great array of location, shape and treatment. Their development includes an extensive chronological variety that exceeds the traditional Breton perspective. In this perspective the use of human figures was exclusively bound to the origin of the construction of the megalithic monuments’ oldest phase. Moreover, the Iberian Peninsula provides references which are not commonly found in the rest of Atlantic Europe: mobile antropomorphic images. These smaller versions help to determine chronologies for the versions on bigger stones.

Thus the name stele suggests a sculptural component, even though its application to the menhir‐stelae in Portugal might be difficult in some cases (Gonçalves et. al., 1997). Its relationship between big stones with rectangular section is countered by big stones with 3 

 

Figure 1.1: Locations of monument with re-used stelae and menhirs in the Iberian Peninsula. circular or oval sections which are called menhirs. Many of the Portuguese menhirs are sculpted and decorated. Therefore there is no doubt about the sculptural intention. On the other hand, there is no real difference between menhir‐stelae and most of the stones from dolmens. Many menhir‐stelae are considered as such because of the open‐air position they hold, either alone or in bigger groups. However, similar stones placed in monuments are called orthostats. The clearest case is the stones which precede the Alberite dolmen in Cádiz: two big stones with rectangular section which probably bore painted decorations. As they are standing on their own outside the monument they are given a different name than the stones situated in the megalithic chamber. The stones, however, are identical (Bueno et. al., 2010).

a synonym of something old and the name menhir‐stele for more recent products (Gomes, 1997; Calado, 2004). The megaliths’ first stones. Ever since the last century’s 90s the relationship between menhirs and megalithic structures has been part of the considerations regarding the origin of Iberian megalithism (Diniz & Calado, 1997; Calado, 1997 & 2002; Gomes, 1996). In accordance with the Breton hypothesis (L’Helgouach, 1983), big stones placed in the open air would have been transported to the first megalithic structures. In this way the initial interpretations of Iberian menhirs, in which they were placed in the final Neolithic, are surpassed (Gomes et. al., 1978).

Probably it is easier to accept these differences than to create new terminologies which would probably only further complicate the matter. In fact, acknowledging this, as we do here, confirms the variability of these expressions ever since their oldest use. Menhirs, stelae, menhir‐stelae and stones from the monuments are different versions of the same stone materialisation system of antropomorphic images (Bueno et. al., 2008; Bueno & Balbín, 1997: 98).

The data to confirm this proposal have reached a complexity which at the time was inconceivable. Several themes from the Peninsula (Bueno et. al., 2007a; Carrera, 2008; Gavilán & Vera, 2005; López et. al., 2009; Oliveira, 2011) have contributed evidences which require detailed analyses. In the current situation it seems to be reasonable to divide our data in two big groups: the ones that certify that stelae and menhirs were the “founders” of megalithic spaces, and the ones that support the idea that, occasionally,

The projection of terminology on the chronology adds more problems. Some researchers use the name menhir as 4 

 

stelae and menhirs were moved in order to create new funerary spaces. “Founding Ancestors” and “Moved Ancestors”, the big stones involved in the construction of Atlantic megaliths, reflect their great ideological importance (Bueno et. al., in press a).

Even considering all of these nuances, we do have cases of “Founding Stones”. The best known examples from the Iberian Peninsula are the defined pits isolated from the other establishment graves of the Casas de D. Pedro dolmen in Cordoba (Gavilán & Vera, 2005). The two menhirs were placed in their original pit. All of the other pieces belonging to the chamber with passage were placed around them later. The two old menhirs are the centre and origin of a funerary structure. A slightly different version is presented by the menhirs from the Granja de Sao Pedro dolmen in Idanha a Velha in the South of Portugal. Their implementation pits were older than the megalith that was built next to them (Almeida & Veiga Ferreira, 1971: 168). As in Casas de D. Pedro, a pair of mehirs is the origin of a posterior funerary installation (Almeida & Veiga Ferreira, 1971: 168).

We are aware of the fact that these words need to be nuanced. It is difficult to include all of the documented objects in one group or the other due to the lack of archaeological information. The great number of reused objects suggests that this possibility should be left open for possible future documentations. Some of the moved stones were only slightly moved from their original position. In fact, the core of the matter is: the fact that anthropomorfic stones are the origin of some monuments equalises both groups in the “founding” role.

Figure 1.2: Pair of founding menhirs of dolmen of Idanha a Velha, South of Portugal and dolmen of Casas de don Pedro, South of Spain . Photo R. de Balbín; Gavilán & Vera, 2005.



 

Recent investigation of the Soto dolmen in Andalusia confirms the use of big menhirs in the chamber’s construction. The pits and their position, together with the headslab which was placed later, confirm the menhirs’ primacy. Moreover, some of the menhirs were reused as a cover for the corridor. This suggests that an older monument was restructured and used to create a new one. Soto is an example of the fact that different situations can be present in the same monument. It is clear that the chamber’s menhirs are founding stones. However, the presence of fragmented stelae in the passage and the discovery of pits outside the monument suggest that some of the pieces present in the monument were transported from another near monument.

The recent work regarding the megalithism of Alter do Châo is a good example. At Anta de Varzea Grande a reused menhir was found, another one at Vale de Carreiras’ dolmen 2 and another one was part of the Anta de Soalheira chamber (Oliveira, 2011). The latter can be identified as a founding menhir, as during the excavations a pit was found that was isolated from the other orthostats around. Several stelae from the Northern Peninsula form the origin of megalithic structures (Carrera, 2008). It’s the same in the South. This is the case of the decorated menhir‐stele from Anta do Telhal in Arraiolos, Evora (Bueno et. al., 2013). It is identical to other stones holding special positions in nearby cromlechs. The Anta do Telhal stele is the first evidence that these sculptures had the same role as menhirs in the generation of new funerary spaces.

In the classical areas of Portuguese megalithism, Alentejo and Algarve, we know of many menhirs that were reused in megaliths. (Calado, 2002; Gomes, 1996). Nonetheless, only few of them have precise archaeological references.

Figure 1.3: Elevational view of stelae and menhir of dolmen of Soto, Huelva, South of Spain. Technical work for Cota Cero SL and pictural’s restitution by the authors.

Figure 1.4: Founding stelae of the anta of Telhal, Arraiolos, Evora, South of Portugal. Photo R. de Balbín.



 

Figure 1.5: Reused menhir of Anta Grande de Zambujeiro, Evora, South of Portugal and a decorated uprigth of its chamber. Photo R. de Balbín. The extension of this situation in the area of Evora, rich in cromlechs and menhir‐stelae, must be even greater. In fact, J. Oliveira has just informed us about the discovery of a destroyed dolmen next to the Evora train station which also contains a menhir‐stele. A bit further to the South, in Proença‐a‐Nova, J. Caninas’ team has found the stele that was part of the Anta do Câo construction.

(Linares, 2011). Another case are the Lagunita I dolmen in Santiago de Alcántara, Cáceres (Bueno, et. al., 2013b). The stones of the oldest monument are completely or partially destroyed. The new monument is then composed of several stelae which we can determine as “re‐ founding” stones. These stones were placed in the pits of the eliminated stones.

We have already documented a reused menhir‐stele of big dimensions at Anta Grande de Zambujero (Bueno & Balbín, 1992: 546). Current work at the site shows that this monument is the sum of several anterior monuments. Founding stones and moved stones are added up in order to create a great structure, dominated by two big stelae (Soares & Tavares, 2010).

The big stones that were moved in order to create new monuments also have, as mentioned above, intermediate situations. Some of them were moved only a little. This is the case of Llano de la Belleza in Huelva (Garcia Sanjuán et. al., 2003). A cromlech underneath the mound is the base of the dolmen that was constructed on top. Some of the cromlech’s menhirs were reused as parts of the dolmen. This type of actions is known in the Iberian Peninsula: the stelae of the Reguers del Seró cist in Lleida come from a nearby line‐up (López, et. al., 2009).

Among the founding stones we need to mention the ones that originate from anterior dolmens that were reconstructed at the same place. As far as we know now this includes several cases in which previous structures were destroyed. This can be deduced from information obtained from the Pozuelo 3 and 4 dolmens in Huelva

In most of the cases we do not know where the reused stones used as pieces, covers or part of the floor in dolmens come from (Bueno et. al., 2007a). Much 7 

 

Figure 1.6: Founding stelae in Antequera megaliths: Cover stone of Viera dolmen, menhir included in the mound of Romeral dolmen, cover stone of Menga dolmen and menhir of Piedras Blancas. archaeological information is missing and, most of all, there are not enough specialists for the study of the objects’ graphics. This is due to the fact that the verification of graphics was carried out after the archaeological interventions. Perhaps the most recent example is brought to us by the Viera dolmen in Málaga. Here we identified the reuse of a stele as a cover for the corridor and the lavish engraved decoration of all of the chamber’s and corridor’s stones after the archaeological fieldwork of the decade of 2000 (Fernández, et. al., 2006). Thanks to the exceptional preservation of the chamber we were also able to confirm the considerable remains of red and black paint on a white base. Four big decorated stelae form a rectangular chamber with a sculpted access. Techniques and manners from the Mediterranean hypogeism are repeated. However, having the archaeological information we do we cannot confirm possible manipulations of the pits, the contemporaneity of the structure or the possible presence of installations associated to the older big stones. This is the option that we proposed for the hill where the Menga and Viera dolmens stand (Bueno, et. al., 2009a: 188).

Geometries of the past Menhirs and stelae that are used as “founding” stones of funerary architectures suggest an anterior phase in which these stones held other positions. This fact, granted by the previously mentioned investigations, has not had the necessary development. This can partially be explained by the lack of data which permit us to confirm structures anterior to the dolmens, as cromlechs and line‐ups present diverse chronologies. Still it has been suggested that circular shapes, named “cromlechs”, are the origin of megalithic architecture. In fact, Portuguese authors (Calado, 1997) have interpreted this type of structures as the origin of megaliths. They base this on the orientation of the circles, which is similar to the orientation of dolmens. Secondly, they base their hypothesis on the tendency of these circles to repeat oval or circular shapes, occasionally in different spaces, as a kind of chambers and corridors. The antiquity of Alentejo contexts would make the region’s circles the oldest antecedents of the Iberian megalithism (Calado & Rocha, 2010). Therefore the circular geometry would be the spatial concept, materialised in the menhir enclosures in



 

interesting point of reflection is the suggestive relationship between ancestral stones that define spaces that have a special meaning and enclosures that repeat the same formulas related with the protection of ancestors.

the South of Portugal, which is the base of all posterior funerary geometries. However, two new elements are added to this hypothesis: firstly the geographical amplitude of the documentation of these cromlechs with older chronologies than other sites in the Peninsula. Secondly, there is evidence that other geometries played a part in these old structures: line‐ups.

There are also other novelties, discovered thanks to the information from Andalusia. There is a possibility that individual or double structures, on the one hand, and structures with a linear development on the other hand were initial geometries that housed the menhirs and stelae that form the origin of the megaliths.

We have already pointed out the presence of old cromlechs in Catalonia (Bueno et. al., 2005 & 2007a). Information from Andalusia has increased remarkably. The Pasada del Abad, Corteganesa and other circles in Huelva (Linares, 2011) emphasise the importance of this dynamic beyond Alentejo. The best reference is the mentioned Llano de la Belleza dolmen in Aroche, Huelva (Garcia Sanjuán et. al., 2003). The area’s geophysical research suggests the presence of a cromlech previous to the dolmen.

Solitary menhirs, such as the one at Alberquilla and many more which we will not mention now, seem to be one of the resources where founding stones were acquired for the construction of megaliths. An inventory of these monuments in Andalusia still needs to be made, but over the past years their documentation has increased noticeably (Linares, 2011; Bueno et. al., 2004 & 2010). Evidences of pairs of menhirs, such as the ones of Casas de Don Pedro and the Granja da Idanha a Velha sepulchres, propose interesting reflections regarding the figures of pairs or twins in European Prehistory (Bueno et. al., in press b; Kristiansen, 2010).

The cromlech underneath the Llano de la Belleza dolmen includes a section of a ditched enclosure that connects some of its menhirs. Therefore Llano de la Belleza proves the existence of old cromlechs in Andalusia and the close connection with the origin of the dolmens’ construction. Moreover, it adds another important factor to our current knowledge. The relationship between the stones of a cromlech and a ditched enclosure signals the ideological connection between the lay‐out of the ditched enclosures and the lay‐out of stone cercles. Cromlechs and ditched enclosures show a series of important concomitances that emphasise the symbolical contents of several structures.

Authentic line‐ups have been found in the North of Portugal (Silva, 2003) and in Catalonia (Lopez et. al., 2009). Both examples can chronologically be placed between the end of the IV millennium and the first half of the III millennium cal BC. This chronology supports our hypothesis stating that the processes of “founding stones” are extended in time and coincide with the diachrony of megaliths (Bueno et. al., 2007a).

The most complex stone structures from Alentejo validate this hypothesis. M. V. Gomes (1997) evaluated the archaeological information of the cromlechs of Almendres and Portela dos Mogos, saying that they were a group that was constructed over time. This idea of a progressive construction of the stone circles fits better with what we know about other European areas (Pollar & Ruggles, 2001). Their connection with the circular growth of important ditched enclosures, such as the nearby and spectacular site of Perdigôes, is very convincing. In fact, A.Valera proposes a possible presence of wooden posts. He also connects the ivory decoration found at Perdigôes with the engravings on many of the menhir‐stelae of the cromlechs of Alentejo (Valera, 2010a). The recent discovery of a wooden post with megalithic engravings in Wales (Jones, 2013) is added to the wooden posts that accompany the stones of the Saint‐Just alignments in Brittany (Le Roux et. al., 1989). All this indicates that the wooden versions had a much more important role than we suspected.

However, some data seem to be older. Although they still lack the necessary chronological precision, the work at Piedras Blancas in Antequera is very promising. The stones, lined up from East to West, among which Piedras Blancas II and III preside, have a collection of lithics with an evident chronology in the Piedras Blancas I cromlechs: blades and microliths (Bueno et. al., 2008, 2009a; Garcia San Juan et. al., 2009). Lined up stones seem to be the origin of the reused stones at the Pozuelo 3 dolmen, according the archaeological information obtained by Linares (2011). It seems convincing for a grouping with this geometry to be connected with the reused stelae from the Soto dolmen, also in Huelva. The variety of these geometries brings the data from the Iberian Peninsula closer to other European cases, especially to Breton data. Line‐ups and cromlechs were the origin of the oldest megaliths. Among all of the data we work with there is a certain tendency to proliferate the Breton line‐ups and the cromlechs from the Iberian Peninsula. This discernment, which is not going to be established rigurously, could have some repercussion on the representativity of circular lay‐outs of Iberian monuments. We do not intend to rekindle the ancient discussion of “circle/line” by this, which has had so much

The antiquity of the ditched enclosures is still being studied. We know that the oldest precinct of Perdigôes might have been created in the Final Neolithic (Valera, 2010b). However, older ditched enclosures must have existed as well. This is suggested by information from the Iberian Peninsula (Rojo et. al., 2008), concording with what has been found at Llano de la Belleza. An 9 

 

All come from sepulchres of varying architectural typologies. This, rules out the old assumption that megalithic decorations are exclusively associated with chambers with corridors (Shee, 1981). But perhaps the greatest novelty of the past years is the verification of reuse in sepulchres of the advanced megalithism. We are referring to sepulchres with false domes, galleries with varied lay‐outs or, in the case of Extremadura, necropoles consisting of small sepulchres which emulate the classical formula of chambers with corridors or chambers without corridors.

impact on the study of ceramic decorations and other European Neolithic matters. On the contrary, circular and linear geometries are part of our European registers during the whole Neolithic and, naturally, before as well. However, this tendency could propose interesting future reflections. Human images as markers of time and space Old menhirs and stelae were the origin of the oldest dolmens. However, also more recent menhirs and stelae were part of the materials used to build dolmens throughout the constructive dynamic of these monuments. The intensification of these processes in the advanced megalithism is exemplified in several cases in Andalusia. There are for example the datings in the Third millennium cal. BC from the Los Gabrieles necropole in Huelva which contains two reused menhirs (Linares, 2011; Linares & Garcia Sanjuan, 2010). And also the post quem reference for the Soto dolmen where reused menhirs were included is a good example.

Stelae that were placed at the entrance of monuments and in visible areas are, as we mentioned above, the materialisation of a classical definition of protective stones. Important case studies are the Palacio III stele from Seville, placed at the entrance of a hypogeum sepulchre and the Toniñuelo stele from Badajoz, placed at the entrance of a chamber with a false dome. The first is an example of a monument that has been excavated recently and provides valid archaeological information. The latter is the opposite. This site was excavated a long time ago and there are severe problems when it comes to establishing the contexts. Nonetheless, both sepulchres from the advanced megalithism show elaborate interior, and probably also exterior decorations. The perception of these funerary complexes in their territory proposes data regarding their roots in older populations and their persistence.

Although it is difficult to establish the chronological aspects it is true that we need to add the direct chronologies from reused painted stelae in megaliths to the datings that were proposed to establish the age of menhirs (Carrera & Fábregas 2002; Bueno et. al., 2007: 600). The previously mentioned datings indicate the same age. Diachrony is one of the most important parts of all of these objects. Even more so, when we have, evidences of their role in the ritual of the megaliths’ construction, which implies physical movement. The reuse of stones and the changing of sepulchres in the IV and III millennium cal BC are some of the most outstanding evidences of the recently acquired documentation from the South of the Iberian Peninsula.

Palacio III contains an older dolmen and a Final Bronze Age tomb (Garcia Sanjuán, 2005) within its mound. Granja de Toniñuelo presents indications of a settlement beneath the mound and evidence of funerary occupations from the Bronze Age and Iron Age. The Roman occupation on top of the mound is the clearest (Carrasco, 2000: 299, 306). Both sepulchres were elaborately decorated and there are evidences of maintenance and remake (Bueno & Balbín, 1997). Paintings, engravings and stelae are evidences of the symbolical complexity of these funerary spaces in which the stelae hold very visible positions. In this sense they are an identification of the monument, apart from being an identification of the people who were buried in it.

Therefore it is now time to question the valuation of menhirs as the only proof of an old phase of megalithism. Moreover, it is time to start working on the perspective of establishing strategies that propose solutions for the anterior structures where these menhirs come from. The custodian stelae would be all of the sepulchre’s components. Latent and patent ancestors, who have received an express decoration (Bueno & Balbín, 1994), accompany the place where ancestors are deposited, adding memories and generating a past to which new memories are also added. Stones with geometrical cloaks reproduce clothing that has been found on mobile objects (Bueno, 2010). The most distinguished ones are placed inside and outside funerary spaces, indicating introductory positions that mark the rythms, times and spaces of the death ritual.

The difficulties of situating the Granja de Toniñuelo stele in relation with the sepulchre start in the 60s. In that moment Almagro focused his work regarding Final Bronze Age stelae on the gathering of all documented objects of similar chronologies in the Southwest (Almagro Bash, 1966). However, the text written by the Leisners is very clear, although some authors have tried to twist its reasoning recently (Cardoso, 2011).

Over the past years references have piled up, including indication menhirs at the entrance of dolmens, necropole indication menhirs, closing menhirs and stelae of some monuments and menhirs and stelae on mounds (Bueno et. al., 2007a, 2010 & 2011).

The relationship between sepulchre and stele is quite convincing, as G. Leisner says in the five pages which he dedicated to this topic. Apart from that, all of the materials that were being kept in the farmhouse come from the surroundings. We personally confirmed that at the end of the 70s. 10 

 

It is important to point out that Toniñuelo is the only one of the Centralwestern stelae which is made from granite and treated as an orthostat. It even has a base of incised elements, made with the same technique as the one applied to the other decorated orthostats in the chamber. This is also visible in our sketch (Bueno & Balbín, 1997: 114).

elaboration of these installations, their ritual administration, and their transformation in stone structures is probably the work of the first farmers. By placing the stones they expressed the perpetuity of their ancestral territories and the indicators that made them recognisable. The first farmers’ symbology and its elaboration in old megalithic sepulchres is not abondened. On the contrary, from the second half of the IV millennium cal BC onwards and above all during the III millennium cal BC it holds an even stronger position (Bueno et. al., 2001 & 2007a). Symbological resources are the basis of the ideological display that is best represented by lineage and inheritance (Bueno et. al., 2005, 2007c).

On the other hand, as Leisner also said (1935: 133), the chronology of the monument throughout the III millennium cal BC is perfectly compatible with the chronology we can appoint to this stele (Bueno, 1995; Bueno et. al., 2005). The association of armed stelae with prestige objects in the Bell Beaker sphere is exemplified by several sepulchres from Andalusia. Amber from Sicily, African ivory, gold and variscite from the Iberian Peninsula are part of a group of exchange materials that is present in very distinguished funerary ostentations throughout the III millennium cal BC. The ideology of these displays is similar to that of Atlantic megalithic art. Especially in the Peninsula it also contains an indisputable base that could have been more important than we thought when it comes to the generation of formulas that are associated to antropomorphic representations in the Mediterranean.

Acknowledgments This paper is part of the results of research project “Graphics programs in Andalucia megaliths” HAR200806140 of the Ministry of Economy and Competitiveness of Spain. We are grateful to our colleges F. Giles, J. Ramos, J.M. Gutierrez, L. Garcia Sanjuán, J. Linares Catela, J.C. Vera, R. Martinez, J. Oliveira and J. Caninas. We are particularly grateful to J.A. Linares to permit the reproduction of reused stones of dolmen of Soto where we work togheter.

This places the Iberian Peninsula in the same spot as the European metallurgic sites which have clear evidences of the exhibition of power using ancestors in the form of the materialisation of armed statuary (Bueno et. al., 2005). Over the next years several issues that are only now starting to be discerned will be resolved.

References ALMAGRO BASCH, M. (1966) – Las estelas decoradas del Suroeste. Biblioteca Praehistórica Hispana. Madrid: CSIC, vol. III. ALMEIDA, F. DE, FERREIRA, O DA V. (1971) – Um monumento pré-histórico na Granja de S. Pedro (Idanha-a-Velha). Actas do II Congresso Nacional de Arqueología. Lisboa: [s.n.], vol. I, p. 163-168. ARNAL, J. (1976) – Les statues-menhirs, hommes et dieux. Paris: Hespérides, p. 239. BUENO RAMÍREZ, P. (1992) – Les plaques decorées alentejaines: approche de leur étude et analyse. L’Anthropologie. Paris: S.A.P, t. 96, nº 2-3, p. 573604. – 1995: Megalitismo, estátuas y estelas en España. Notizie Archeologiche Bergomensi. Bergamo, 3, p. 77-129. – 2010: Ancestros e imágenes antropomorfas muebles en el ámbito del megalitismo occidental: las placas decoradas. In CACHO, C.; MAICAS, R.; GALÁN, E. and MARTOS, J.A. (coord.) – Ojos que nunca se cierran. Ídolos en las primeras sociedades campesinas. Ministerio de Cultura. BRADLEY, R. (2009) – Image and Audience: Rethinking Prehistoric Art. Oxford: Oxford University Press. BUENO, P. and BALBIN, R. de (1992) – L’ Art mégalithique dans la Péninsule Ibérique. Une vie d’ensemble. L’Anthropologie. Paris: S.A.P, t. 96, nº 2-3, p. 499-570. – (1994) – Estatuas-menhir y estelas antropomorfas en megalitos ibéricos. Una hipótesis de interpretación del espacio funerario. Homenaje al prof. Echegaray.

The beginning of megaliths During the past years the interpretation of European megalithism has improved drastically. This is partially thanks to finer analyses of the movements of the stones used for these structures. It is generally admitted that sepulchres are the final product of remake and maintenance (Laporte, 2010). We know that their origin lies with ancestral stones. Probably the clearest antecedents of these stones are wooden representations. There are starting to be several examples of the famous wooden menhirs, usually found at Mesolithic settlements in the North of Europe, at other sites of the Atlantic façade (Rust, 1943). Although we are still in need of more detailed studies two lines of work stand out: firstly, their chronologies are very similar, or even identical to the oldest direct datings of painted stelae in the North of the Iberian Peninsula (Carrera & Fábregas, 2006). Secondly, the decorative techniques originate from the European hunting population to a great extent (Bueno et. al., 2007b). The placing of wooden posts, probably including several decorated objects, pit circles and line‐ups must have been part of the social cohesion systems and the definition of the use of the territory from the Mesolithic onwards. The 11 

 

C14 et contextes archéologiques. L’Anthropologie. Paris: Elsevier, t. 111, nº 4, p. 590-654. – (2007c) – Ideología de los primeros agricultores en el Sur de Europa: las más antiguas cronologías del arte megalítico ibérico. Cuadernos de Arte Rupestre. Moratalla: Centro de Interpretación de Arte Rupestre de Moratalla, 4, p. 281-312. – (2008) – Dioses y antepasados que salen de las piedras. Boletín del Instituto Andaluz del Patrimonio Histórico. Sevilla: Instituto Andaluz del Patrimonio Histórico, vol. 67, p. 62-67. – (2009a) – Análisis de las grafías megalíticas de los dólmenes de Antequera y su entorno. Dólmenes de Antequera: tutela y valorización hoy. Sevilla: Instituto Andaluz del Patrimonio Histórico, PH cuadernos, nº 23, p. 186-197. – (In press a) – Graphic programs as ideological construction of the megaliths: the south of the Iberian Peninsula as case study. IIº Congresso Internacional sobre Arqueologia de Transição: O Mundo Funerário. Évora, Portugal (29-30 abril 2013). – (In press b) – Human images, images of ancestors, identitary images. The South of the Iberian Peninsula. 3e Colloque international sur les pierres levées et statues menhirs au Néolithique. Saint-Ponsde-Thomières.

Altamira: Museo y Centro de Altamira, Monografías, 17, p. 337-347. – (1997) – Arte megalítico en sepulcros de falsa cúpula. A propósito del monumento de Granja de Toniñuelo (Badajoz). Brigantium. La Coruña, Spain: Museo Arqueolóxico e Histórico, 10, p. 91-122. – (2006) – Arte megalítico en la Península Ibérica: contextos materiales y simbólicos para el arte esquemático. In MARTÍNEZ, J., HERNÁNDEZ, M. (eds.) – Arte rupestre Esquemático en la Península Ibérica. Comarca de Los Vélez, p. 57-84. BUENO, P., BALBIN, R. and ALCOLEA, J. (2007b) – Style V dans le bassin du Douro. Tradition et changement dans les graphies des chasseurs du Paléolithique Supérieur européen. L’Anthropologie. Paris: Elsevier, t. 111, nº 4, p. 549-589. BUENO, P., BALBÍN, R. DE. AND BARROSO, R. (2004) – Arte Megalítico en Andalucía: una propuesta para su valoración global en el ámbito de las grafías de los conjuntos productores del Sur de Europa. Mainake. Málaga: Centro de Ediciones de la Diputación de Málaga (CEDMA), XXVI, p. 29-62. – (2005) – Hierarchisation et métallurgie: statues armées dans la Péninsule Ibérique. L’Anthropologie. Paris: Elsevier, t. 109, nº 4, p. 577-640. – (2007a) – Chronologie de l’art megalithique ibérique:

12 

 

The contribution of Manuel Heleno to the knowledge of the funerary Megalithic in Alentejo Leonor Rocha

part of the Geological Commission, Carlos Ribeiro, Nery Delgado and F. Pereira da Costa, Estacio da Veiga, Santos Rocha, among others.

Abstract The excavations conducted by Prof. Manuel Heleno in funerary megalithic monuments in the Central Alentejo, in the 30’s, during the twentieth century, were considered for decades the key to the knowledge of its evolution. This mythical idea naturally originated from the fact that this researcher never published his work.In fact, a review of his fieldwork, conducted recently by the signatory, helped confirm that despite the invaluable contribution that his research represents, the data collected did not unveil the genesis and evolution of the megalithic architecture.

Leite de Vasconcellos stands out for being the precursor of a national project that spurred the Portuguese archaeology, as founder and director of the Portuguese Ethnological Museum. In fact, the project of bringing together, at that Museum, a representative collection of the whole national territory, led him to establish a network of regional informants that in a way, helped to boost the regional archaeology and to make it known. Manuel Heleno is, in this perspective, a faithful follower of this policy, substantially expanding his action, which often resulted in criticism and controversy within the scientific community. Three factors contributed, in my view, to this situation: his age, his rapid career growth and the “protection” given to him by the existing legislative framework at the time.

With this work we seek to present his main ideas by comparing them with the results of more recent studies, based on work performed by the signatory and by other researchers. Keywords: Manuel Alentejo; Evolution

Heleno;

Funerary

Megalithic; In terms of age, it is not now, nor was it then, common for such a young person (35 years old) to be the Director of a National Museum and become an Archaeology Professor at the Faculty Letters of the Lisbon University (39 years old). As I mentioned earlier, obtaining almost unlimited management power of the national archaeology, by virtue of his position as Director of the Museum, brought him various kinds of problems.

Manuel Heleno: polemics of an archaeologist The second half of the nineteenth century marks an inaugural step in the Portuguese archaeology, reflecting, possibly with some delay, the global progress of the matter. In Portugal, this process is, to some extent, relatable to the new mind-set in the wake of the political changes of 1834, which allowed the development of a “modern scientific and technological spirit”, which was reflected in the cultural field, through the creation of the Thought and Action infrastructures “(Diniz & Gonçalves, 1993-1994: 179).

According to J. L. Cardoso, “that institution held authority to intervene in archaeological findings that took place anywhere in the country, which, of course, could create friction with local researchers or other Institutions that had legitimate priority over the discoveries” (Cardoso, 1993 -1994: 298).

It is a period of great cultural dynamics, with the appearance of a vast movement of regional and local historiography, of positivist level, which intended to create the basis for the History of Portugal, on scientific grounds. In a more or less general manner, historians interested in local and regional history appeared throughout the country, committed to basing their investigations on sources that were archival, archaeological, ethnographic, etc. This new impetus led to the creation of several societies, related to natural and social sciences that were reflected in the creation of multiple scientific periodical publications, and the emergence of a group of researchers who distinguished themselves bythe quality of their work.

In fact, the publishing of Decree 21117, April 18th, 1932, Excavations and Lien/Listing of national antiquities, which regulated the archaeological activity in Portugal (Chapter III), gave the Director of the Museum the power to authorize, supervise and suspend the archaeological excavations that took place in Portugal, and also give it the possibility to claim for itself the scientific priority over sites considered most relevant. This legislation became the basis of all conflicts that began as early as the following year, between Manuel Heleno and other archaeologists, Mendes Corrêa in particular. In fact, 1933 was fruitful in conflicts, between these two archaeologists, that came to be of public domain through the published letters in the daily press, with accusations from both sides. The emphasis behind the criticisms that were made of him where, precisely, his lack of experience and his arrogance.

Among these Portuguese archaeology pioneers, it is fair to name a few of those that substantially contributed to the advancement and consolidation of prehistoric archaeology in Portugal, especially researchers that were 13 

 

Map 2.1: study area Going back to the events and existing data (Rocha, 2005) we can see that Manuel Heleno trusted no one: not his most direct employees, and not even in researchers who shared some fieldwork with him, such as Leite de Vasconcelos and the Leisner couple. During the excavation of the passage tombs at Herdade Grande, which took place under the direction of Leite de Vasconcelos, we can see that the numbers assigned by Manuel Heleno in his Field Notebooks (Rocha, 2005) do not correspond to those of Leite de Vasconcelos (Vasconcellos, 1929: 160-169). Also in the excavation of Vale das Covas (Coruche) passage grave, which he shared with the Leisner couple and Obermaier, Manuel Heleno changed the names of the monument, seeing that the Leisners refer another name (Leisner & Leisner 1959: 274).

Effectively, the work on the megalithic he came to develop in the Central Alentejo for about two decades (Map 2.1), showed his weaknesses and especially his inability to communicate with his peers. Despite having recorded and studied about three hundred megalithic monuments, Manuel Heleno did not publish any of this information, thus depriving the scientific community of data that could have allowed research development. However, it was this attitude that eventually created and fed a myth around the legendary monuments that had the “key” to decode the origins and evolution of the megalithic phenomenon in this region, seeing that Manuel Heleno only revealed the alleged results of his studies via daily press and in some lectures, as part of his activity as a Professor at the Faculdade de Letras of Lisbon.

This “shuffling” of information, so that no one would be able to understand his research, seems to be a constant in Manuel Heleno’s life, reason why he also chose to encode all megalithic monuments he studied, writing

Without scientific publications, without presenting results, without showing archaeological findings, everyone was limited to mere assumptions based on comments made by Manuel Heleno. 14 

 

down the codes in his also mythical Field Notebooks, which he kept at all times.

shell mounds burials, having been done, afterwards, individually, in graves “(Rocha, 2005 Cd.14 - Volume 2, Annex 1, p. 101). However, the excavations of the Muge shell mounds were the responsibility of Rui Serpa Pinto, Correia Mendes’ staff member, who had just done a first systematic study of its lithic industries (Pinto, 1932), a study to which Manuel Helenus obviously had access to. The different perspectives on the origins of the Portuguese race defended by Manuel Heleno and Mendes Correia and which were mostly based on anthropological remains and existing industries in the Tagus, were also at stake. In fact, Mendes Correia argued that the Muge shell mounds had their genetic roots in North African populations, but that the current Portuguese population originated from the dolmen builders, and that there is no genetic link between the two groups (Correia, 1919, 1923, 1938). Now this was an idea that Manuel Heleno tried to refute for considering the two proposed theories unreasonable: on the one hand, the eventual negroid character of the Tagus Mesolithic, seeing that “recent research” does not authorize stating “the African origin of these industries” and would support instead “an European affiliation of our more remote ancestor” (Heleno, 1944: 8). On the other hand, the lack of a continuity line between the Muge Mesolithic populations and the builders of the first megalithic tombs, as he advocates, regarding the origin of the small graves “these were created by the “capsenses”, as evidenced by the silica analogous to the ones at Mugem, which has earlier burials in the shell mounds, and then having started doing them individually in graves “(Rocha, 2005 Cd.14 Volume 2, Annex 1, p. 101). Without access to the Muge excavations and in order to seek to counter Mendes Correia’s ideas, Manuel Heleno then turns to the Sado...

From the idea to the data: the search for what is real in Manuel Heleno’s imaginary The model outlined by Manuel Heleno for the origins and evolution of the Alentejo megaliths is based on an indigenous perspective (Heleno, 1956), allegedly based on excavations done during the 30’s and 40’s, of the twentieth century, in the Central Alentejo. This approach integrated what was ideal at the time, whereby it was sought to unveil, based on archaeological data, the very origins of the “Portuguese nation” (Fabião, 1999). As we know, the nationalism that flourished at the time sought, above all, “to reinforce the pride and morale of nations and ethnic groups” (Trigger, 1992: 167) based on an archaeological discourse, whereby one of the most notorious and extreme examples was that of Gustaf Kossina, in Nazi Germany. The position of Manuel Heleno, in relation to the evolutionary sequence of the Alentejo megalithic funerary, is summarized in one of his Field Notebooks (Cardoso, 2002: 188; Rocha, 2005, Volume 2, CD. 28, Annex 1, p. 183) where he presents his more general conclusions: the oldest graves were the small closed ones, and the most recent, the chambered ones with a long passage and lobby. Manuel Heleno’s major concern focuses in fact on this issue: the origins and evolution of funerary megalithic. His Field Notebooks present, from the beginning, scattered summaries and brief comments, that reveal his concern in confirming, on the one hand, and to develop, on the other, an evolutionary model that, in broad terms, had, most probably, been conceived even before the work began. Although he never acknowledged it, the fact that his boundaries coincide with those of Virgil Correia’s (Correia, 1921), the fact that he excavated again some of Brisso’s set of monuments (Pavia, Mora) and that in Pavia there were small graves and large passage graves, leads me to believe that Manuel Heleno conceived his project and the megalithic interpretive scheme based on V. Correia’s data that, as we know, was deposited in the Ethnological Museum, as Leite Vasconcellos withheld this information, not allowing V. Correia to use it in his publication.

In fact, the excavations conducted by Manuel Heleno in megalithic monuments gave him contradictory data that, for various kinds of reasons - perhaps including his lack of scientific basis on this subject and his excess work load (Museum Director, Professor at the Faculdade de Letras and the management of numerous archaeological sites) - he was unable to resolve. In fact, as the data did not fit into the previously formulated theories, he tried to redesign his interpretive schemes: In 1936, in Notebook 28, based on the results obtained at the Herdade do Deserto (Montemor-o-Novo) Manuel Heleno concludes that “the passage graves’ architecture evolvedas follows: I – box-shaped passage graves, circular, with little or no potteries, semi-lunar silica and trapezoidal or triangular; II – passage graves of the same type but elongated, strangled or divided into two compartments (collective). Alongside this the smaller and more antique types.

This search for origins, which has advances and retrogressions along his 45 Field Notebooks, depending on the data that he collects in his excavations, has other setbacks, resulting from some elements that Manuel Heleno would have liked to have controlled: access to other archaeological sites, particularly the Tagus shell mounds.

From these onwards, a double evolution: A) The large graves move on to have large chambers and passage; B) The small graves, with only a “chamber, which extends, and under the influence of the huts, starts

In fact, in the first years, Manuel Heleno states, regarding the origin of the small graves, that “these have been created by “capsenses”. This is evidenced by the silica similar to those at Mugem, which were at the earliest 15 

 

having a round shape, and is of the same type that evolve into the ones with a chamber and round lobby.” (Rocha, 2005, Cd. 28 - Volume 2, Annex 1, p. 183).

and from North to South, while, on the contrary, the schist plaques increased in opposite directions. In addition to these summaries that he completes, there are other comments that are important to record. At the end of the description of each of the excavation monuments, Manuel Heleno usually made a brief comment on what could be on any of the findings that he considered important, namely on the excavation, on the architecture, etc.. However, there is a rank assigned to the architecture of some monuments, the small “primitive” graves, which appear ranked by “A”, “B” and “C” (Ibid., Cd 2 - Volume 2, Annex 1, p. 268 ff.) the meaning of which he never explains. The confrontation of these notes with the analysis of the description of the monuments, leads me to conclude that within the group that he names “primitive”, he considered a classification based on the geometry of the plans, which is systematized in the following table.

In 1937, in Notebook 31, he reshapes the sequence by removing the link between the architecture and findings for the oldest, considering only that: “evolution: – Passage graves with gallery; – Passage graves with two compartments in the shape of gallery, trapezoidal or circular; – Passage graves with chamber and passagewith pillars at almost the same height; – Passage graves with chamber and wide passage; – Passage graves with a round chamber and round lobby” (Ibid., Cd.31 - Volume 2, Annex 1, p. 207). Still in 1937, in Notebook 32, he reflects on the connection between the architectures and the findings “in the classification of dolmens it is necessary to consider not only the architecture, but also the evolution of the material. There are indeed primitive architectural forms in later periods/surviving material (archaic) in modern passage graves” (Ibid., Cd.32 - Volume 2, Annex 1, p. 214), concluding that: “To our current knowledge, resulting from the 1930 excavations, we can make the following classification: GROUP I - semi-spherical pots a) Box-shaped passage graves without pottery and with silica and axes. b) Box-shaped passage graves with pottery, silica and no axes. c) Box-shaped passage graves with no pottery and no axes.

PRIMITIVE GRAVES

Although, for the passage graves, which he often designates as “evolved”, he did not code the megalithic architectures, an analysis of his speech allowed me to distinguish three major groups.

EVOLVED PASSAGE GRAVES

Circular chamber and passage grave Large chamber and long passage Chamber, passage and lobby

However, what interfered most with Manuel Heleno’s idealized model was the combination of the findings with the megalithic architectures. In fact, if in some monuments, he found what he considered to be chronologically fitting, there were others that disturbed the sequentially established order. In a way, these differences obliged him to a volte face... “in a primitive way. However the axes are already very polished.” (Idem, Cd.18 – Volume 2, Annex 2, p. 126).

GROUP II - passage graves with silica with dent, a lot of pottery and axes. a) Gallery shape, due to the fact that they have become collective graves. b) Passage graves, halved, in compartments, preceded by elongated, strangled. GROUP III - influenced by the round shape of the huts, the following derived: – From GROUP I the passage graves only with a round chamber, and by transition, with the oval shape – From GROUP II the passage graves with chamber and passage, with the 1st phase with the Oliveira, Bertiando sand Pasmaceira types of Cross. These types moved on to the chamber and round gallery type, Rabaçal type with predominance of convex based arrows and anthropomorphic schist plaques.” (Ibid., Cd.32 - Volume 2, Annex 1, p. 213)

The polished stone axes were undoubtedly one of the things that most troubled him. He took note, expressly, with some perplexity that, for example, at passage grave V “five axes appeared (...) of various kinds: a poorly polished one, rectangular; two other round ones; two flat ones. It is worth noting the difference between the types of axes from the same time period.” (Ibid., Cd.9 Volume 2, Annex 2, p. 71), or even, in regard to the ER passage grave, he stresses the “association between the round axes and the sub-square” (Ibid., Cd.28 - Volume 2, Annex 1, p. 184).

Already in 1939 in Notebook 39 (Ibid., Volume 2, Annex 1, p. 253), Manuel Heleno chooses to sketch a new space model, whereby he compares the data collected in the area he studied, considering that one could observe an increase in the size of the monuments, from West to East

Axes with a “rough appearance” poorly polished or perforated, square or round, are in his point of view the most antique, but the issue of the polishing of axes had evolved. In fact, the initial theory that considered the lack 16 

 

A: Rectangle B: Trapezoidal C: Trapezoidal Circular B/C + incipient passage

Huelva (Ibid., Tafel, 45, 3-5; Tafel, 46, 2, 3, 7, 11; Tafel, 48, 1-4). This type of geometric, in contrast, is totally absent; it seems, from the findings recovered at the monuments of the district of Portalegre (Oliveira, 1998b).

of polish to be synonymous with ancient is abandoned, from 1939 onwards, concluding that this appearance was not a good chronological indicator. Regarding the chronology of the polished stone axes, I believe that today it is relatively clear that there is precedence to the axes with the rounded cross-section, the general shape that is obtained by perforating or polishing, in relation to the polygonal cross-section ones, which usually generally resulted from the longitudinal chipping. In this regard, the observations made by Manuel Heleno were correct, and were merely contaminated by criteria which appeared to be irrelevant, as is the one referring to the extension of the polishing and its co-existence with the other types.

However, it is the extraordinary survival of the geometric, throughout the megalithic sequence, that naturally caused the greatest perplexities and confused the interpretation. Manuel Heleno refers to the existence of various geometrics in several megalithic monuments with passage, such as the A do Paço passage grave, where he will have collected microliths in a hole in the pillar; this stratigraphic observation may, eventually, allow the interpretation of the deposition of these artefacts, in a complex monument, as part of their foundation ritual, therefor previous to the beginning of any primary depositions (Rocha, 2005).

The Chipped Stone on the other hand, especially the Geometric Group, are, undoubtedly, his preferred elements that allowed him to make many chronological comments. For the geometric in general, a Mesolithic origin is repeatedly invoked, surely having in mind the lithic industries of the Tagus shell mounds. Regarding the specific evolution of the geometric, Manuel Heleno states in 1937, in Notebook 32 (Ibid., Volume 2, Annex 1, p. 214), that “in the silica the evolution is more certain: 1st Semicircle; 2nd Trapezoidal; 3rd Trapezoidal or semicircle with dent; 4th Triangular already with a concave base “fitting chronologically into the Neolithic.”

In terms of recorded internal variability, regarding the geometric present in megalithic monuments, it can be seen that, in general, the trapezoids predominate, and, within these, the asymmetric, followed by the increasing, while the triangles are, in most cases, underrepresented. This sequence shows some differences in comparison to the Mesolithic findings and, in particular, in relation to the Moita do Sebastião, one of most antique shell mounds. The explanations for these discrepancies have been based around relatively vague concepts - survival and archaisms - for which there is no independent confirmation, or even attempts to deny the obvious contemporary relationships (Andrés Rupérez, 2000: 262).

Initially, when microliths arise in later settings, in other words, in typological more complex and larger monuments, Manuel Heleno suggests that it is a “Mesolithic survival”. However, the repeated presence of such findings in larger monuments initially leads him to the conclusion that these “lasted until the Early Eolithic” (Ibid., Cd.7, Passage Grave O - Volume 2, Appendix 1, p. 47) and, later, considered that they had a long diachrony, and said that “in any case the silica are largely Chalcolithic” (Ibid., Cd39, Passage grave KX - Volume 2, Annex 1, p. 251).

Regarding the arrowheads, Manuel Heleno seeks, when searching for a linear typological sequence that characterizes most of his proposals, to explain the appearance of concave based arrowheads; this is a result of an evolution from geometric (triangles or trapezoids ) with a concave truncation, which he classifies as “concaved based silica”. It is noted that, on the basis of the same type of approach, the Leisners suggest precisely the reverse, assigning the trapezoids “with a slightly concave bottom side” to the “influence of the concave based retouched arrowhead” (Leisner, 1951: 54).

Note that in the proposed sequencing reproduced above, the silica, besides appearing isolated, as in a final stage of the process, the triangular silica with lower concave truncation - which, as we have seen, announce the concave based arrowheads - also highlight the “dented” silica (16 monuments).

As for the “convex based arrows”, as Manuel Heleno could not find a credible background for them in the geometric, he considered them to be “of foreign origin and having been predominant in the first phase of the Chalcolithic”. In the same view, he also proposes that “the dagger and the halberd result from these” (Rocha, 2005, Cd.2 of Estremoz – Volume 2, Annex 1, p. 269).

After him, not enough attention has been given to this by archaeological literature, yet it has been isolated by the Leisners as one of the present types (although only 2 samples) in Reguengos de Monsaraz (Leisner & Leisner 1985, 56). Besides these two samples, the German authors published another doubtful sample from the Capela Passage Grave (Avis) (Leisner & Leisner, 1959: Tafel 15, 3, nº 23) and, surprisingly, a statistically significant number that originated from some controversial cystoid monuments at Monchique, as well as the multiple elongated chambers at El Pozuelo, in

Sometimes he contrasts the associations of these artefacts with the type of architecture, like in the Nª. Sª. Conceição dos Olivais Passage grave, that “despite its large size, resembling passage graves from the Later Eolithic or even Bronze, it presents concave based silica, similar to those at Oliveira da Cruz, the pair of arrows with a slightly concave base that little exceeds the silica in numbers. “And questions: “is this passage grave antique? (...) In any case it seems that the silica can be kept here longer or passage graves evolved faster. Convex arrows 17 

 

just 10 years (Map 2.2). If we do not take into account the issue of the archaeological recording done in his interventions and the lack of publications, we can consider that part of his work was extremely valuable and that contributed overall to the knowledge of the megaliths. He had against him, at the time, the lack of data from other areas (with exception to V. Correia’s works, in Pavia), which could have been used to compare/assess the results.

maybe from abroad” (Ibid., 2005 Cd.2 Estremoz Volume 2, Annex 1, p. 270). Interestingly, in the latter case, the attempt to explain the observed discrepancies, based on different rhythms of distinct communities, an explanation that, nowadays, still deserves serious consideration, Manuel Heleno does not abandon his “parochial view of developments in closed environments “(Gonçalves, 1992: 104).

In general and taking into account the proposals made by Manuel Heleno; we can conclude that some of them may still be considered generally valid: a) The idea that Alentejo’s funerary megalithic evolved from the simplest to the most complex, whereby, firstly, the graves were closed, followed by small galleries, the galleries divided in half, passage graves that only had a chamber and then the passage graves with a chamber and a passage, ending the sequence in passage graves with long passages, particularly those with a passage that had a wider central area and those in which the presence of a lobby was confirmed. b) The note about the evolution of findings, in particular the polished stone axes, the presence/absence of potteries and geometric. c) Monuments of simpler architecture were associated with the simpler findings with round/oval and square axes and, made of chips with a pierced body or unpolished, geometric and sparse potteries. Others seem to be more debatable as in reality we never had any specific studies in Alentejo, such as: a) The increasing size of the monuments, from West to East and from North to South. The pillars of the monuments that where higher on the East than on the West. On the other hand, the first showing of the tendency for the formation of the passage would be by the difference (decrease), in height, of the final pillars in the elongated graves. b) The schist plaques increased in opposite directions, from East to West.

As we saw before, the arrowheads are clearly interpreted by Manuel Heleno as a late phenomenon, based, in particular, on the associations with materials such as schist plaques, and also based on their relative scarcity in the monuments that he considers more antique. The Schist Plaque Group also raises his interest and so he makes typological and chronological comments along the line; the absence of these artefacts remits the monuments, normally, to older settings, while, in contrast, its presence, places them in the evolved phase (Rocha, 2005 Cd.7, Passage grave O, Volume 2, Annex 1, p. 48). Relatively to the presence of anthropomorphic schist plaques at some monuments, Manuel Heleno believes, in the early stages of his work, such as archaic, particularly when he refers, for the purpose of “anthropomorphic idols” collected at Talha 1 Passage grave (Ibid., Cd. 1 Estremoz - Volume 2, Annex 1, p. 264), that in one of them, the arms appear to “still be drawn with naturalness” and later, for example, in regards to the Freixa passage grave (Ibid., KU Passage grave, Cd 38 - Volume 2, Annex 1, p. 249) and Rabaçal (Ibid., DQ passage grave, Cd.23 - Volume 2, Annex 1, p. 156), he relates an architectural detail that he considers tardy - an antechamber - with the presence of anthropomorphic plaques. The presence of crosiers refers, always, the monument to the Chalcolithic. It is curious, in this regard, the description that he makes of Passage grave 2 of Lobeira de Baixo (Ibid., CI dolmen, Cd 20, Volume 2, Annex 1, p. 129), whereby he refers to a “skeleton (...) with a crosier over its chest” that “it seemed to grip it with the right hand”.

Regarding the evolution of megalithic architectures, in my opinion, the data provided by the excavations conducted post-Manuel Heleno; do not contradict the theory of evolution from the simple to the complex. I would say that this is a general rule of human evolution, in all aspects. As I defended previously (Rocha, 2005) what in reality confused the data and became an insurmountable barrier for Manuel Heleno, was the combination of the findings with the architectures.

As we have seen from previously mentioned examples, Manuel Heleno was, over the years, presenting different evolutionary models. From the idea to the data... this was undoubtedly one of Manuel Heleno’s major handicaps: the model he idealized, from the linear evolution of architectures and findings, had too many discrepancies when compared to the collected data. One thing was what he imagined; the other was the reality of facts.

In fact, while a megalithic structure may be partially remodelled after being built or, ultimately, destroyed, the findings are movable artefacts that can, after having been made, be used and re-used. Therefore, their presence and/or absence may actually have different readings, as we shall see further ahead:

From the previous data to the recent one: the real contribution of Manuel Heleno’s work In terms of the Alentejo funerary megalithic research, when we look at our recent past, we can see that Manuel Heleno accomplished a remarkable job in terms of recording and excavation of megalithic monuments, as until now, no one has been able to match his accomplishments... about three hundred monuments in

In terms of architecture, the data analysis allows us now to see that in fact the area studied by Manuel Heleno has certain uniqueness when compared with the rest of the 18 

 

refer the “polished stone axes + geometric” set, to more recent dates, than those that had previously been proposed. Of course we can always equate, due to lack of absolute dating for the northern part, which small megalithic graves can be previous to these hypogea, and generallyfall into in the 1st half of the fourth millennium BC, or even the end of the 2nd half of the fifth millennium BC.

Alentejo. This specificity is manifested, for example, in the relative abundance of small megalithic graves, which has no parallel in the Alentejo region, which seems to tend to get concentrated further West, decreasing to the East where the passage graves account, virtually, for almost the whole of the monuments, just like in the Reguengos de Monsaraz megalithic set. Manuel Heleno tried to explain this fact on the basis of the geographical proximity to the Tagus Bell Beaker territory, directly where the first builders of the simpler graves would be from. To reinforce this affiliation, we would have the more archaic artefacts, represented by the geometric microliths and polished stone axes, with a circular part and no full polishing. I must highlight that also in the municipality of Montemor-o-Novo, in its SW limit, in the geographical proximity to the Sado shell mounds, a set of small megalithic graves were recently identified (Calado, 2003) which are similar to those identified by Manuel Heleno on the NW side of the municipality (Rocha, 2005).

In fact the question of the evolution of the funerary megalithic, rather than having been solved by Manuel Heleno, as the referenced database was being extended, it was in fact multiplying due to the lack of connection between the architecture and the findings. In subsequent years, the radiocarbon dates and the more attentive stratigraphic analyses came to amplify this problem with the registering of materials allegedly more antique in monuments that were architecturally evolved. On balance, in this final analysis, whereby I tried to analyse the data collected by Manuel Heleno and the most recent data, I presented in 2005 (Rocha, 2005) a proposal for an interpretative model that intended to integrate, in general terms, these alleged anomalies.

The excavations of monuments with similar simple architectures, in the Alentejo (Central and North) in recent years, continue to provide, in general, data that attests the use and re-use of these small graves, in later periods, but which continue to provide the absence of elements (orteological remains, charcoal) that allow absolute dating (Rocha, 2002, 2007b; Rocha and Alvim, 2012, 2013).

In fact, the prolonged use of the necropolis always creates interpretive problems, especially because, if in some periods we find that the new burials seek to respect the previous burial space (Bronze and 1st Iron Age), in others there is an emptying of the entire space, with the removal of all the funerary findings (Roman and later). But beyond that, we can still assume that the materials and osteological remains may have been moved by human groups to other monuments, included in a ritualization phenomenon of new spaces or, simply, to bring the ancestors closer to the new habitat areas. Findings that are removed and/or destroyed by subsequent reuses, findings that are moved to more recent monuments may justify, per si, the observed discrepancy. But we cannot fail to assume the probable existence of some inertia towards innovation, taken at different speeds: for example, at the time the first passage graves with passage were made, maybe the ultimate megalithic tombs, by hypothesis, were carried on being built for some time.

However the situation in the South seems to be different. In fact, the archaeological work conducted in the context of Minimizing Environmental Impacts has revealed a different reality, and the case of Sobreira de Cima seems to be (within the limited published information) the most paradigmatic. It is, on the one hand, another type of funerary monument (hypogea), of which we were unaware of, within the inland of southern Portugal, but that brought a new perspective on the Neolithic burials. The Hypogeum 1, which was preserved with the bones in their correct anatomical position, also had a scant number of burials without overlaps or reuse, and the materials collected in this set provide a total absence of potteries and the presence of geometric and polished stone axes. Of the 5 datings performed, 4 place the use of this necropolis in the last quarter of the fourth millennium BC (between 3240 and 3100 cal BC) (Dias, et. al., 2008; Valera, et. al., 2008). If, on the one hand, these datings confirm the construction and use of such monuments in the 2nd half of the fourth millennium BC, similar to those present in the Central coast/Southern Portugal (Gonçalves, 2010; Soares, 2003), on the other hand they

In short, the evolution of the funerary megalithic appears to be based not on one, but in many evolutionary rhythms that will have to be addressed and reassessed whenever the existence of new data so requests.

19 

 

Map 2.2: Manuel Heleno dolmens de Arqueologia e Património. Lisboa: NIA-ERA Arqueologia, n 2, p. 31-40. DINIZ, M; GONÇALVES, V.S. (1993-1994) – Na 2ª metade do séc. XIX: luzes e sombras sobre a institucionalização da Arqueologia em Portugal. O Arqueólogo Português. Lisboa: Museu Nacional de Arqueologia, Serie IV, nº 11/12, p. 175-187. FABIÃO, C. (1999) – Um século de arqueologia em Portugal – I. Al-madan. Almada: [s.l], II série 8, p. 104-126. GONÇALVES, G. (2010) – Datações de radiocarbono para a Pré-história e Proto-história em Portugal: revisão e calibração. Évora: Universidade de Évora, master dissertation, manuscript. HELENO, M. (1942) – O culto do machado no Calcolítico português. Ethnos. Lisboa: [s.n.], II, p. 461-464. HELENO, M. (1956) – Um quarto de século de investigação arqueológica. O Arqueólogo Português. Lisboa: [s.n.], (n.s.): III, p. 221-237. LEISNER, G. e V. (1959) – Die Megalithgraber der Iberischen Halbinsel: Der Westen. Berlin: Walter de Gruyter. LEISNER, G. e V. (1985) – Antas do concelho de Reguengos de Monsaraz. Lisboa: UNIARCH (reed.). LEISNER, V. (1970) – Micrólitos de tipo tardenoisense em dólmens portugueses. Actas das I Jornadas Arqueológicas da Associação dos Arqueólogos Portugueses. Lisboa: Associação dos Arqueólogos Portugueses, volume II, p. 195-198. OLIVEIRA, J. (1998b) – Monumentos Megalíticos da Bacia Hidrográfica do Rio Sever. Lisboa: Ed. Colibri. ROCHA, L. (2002) – A anta do Couto dos Algarves 2 (Crato). O Arqueólogo Português. Lisboa: MNA, série IV, nº 20, p. 39-60. ROCHA, L. (2005) - Origens do megalitismo funerário

References ANDRES RUPÉREZ, M. (2000) – El Megalitismo en la Cuenca Alta y Media del Ebro. Actas do 3º Congresso de Arqueologia Peninsular. Porto: Adecap, p. 255-269. CALADO, M. (2003) – Megalitismo, megalitismos: o conjunto neolítico do Tojal (Montemor-o-Novo). Trabalhos de Arqueologia. Muita gente, poucas antas? Origens, espaços e contextos do Megalitismo. Lisboa: IPA, Actas do II Colóquio Internacional sobre Megalitismo, nº 25, p. 351-369. CARDOSO, J. L. (1993 – 1994) – A arqueologia portuguesa do pós-guerra vista pela correspondência de O. Da Veiga Ferreira e Abel Viana. O Arqueólogo Português. Lisboa: Museu Nacional de Arqueologia, Serie IV, nº 11/12, p. 291-338. CORREIA, A.A. Mendes (1919) – Origins of the Portuguese. American Journal of Physical Anthropology (Am J Phys Anthropol.). USA: American Association of Physical Anthropologists, volume 2, nº 2, p. 117-145. CORREIA, A. A. Mendes (1923) – Nouvelles observations sur “l`Homo taganus, Nob.”. Revue Anthropologique. Paris: Librairie Émile Nourri, A33, nº 11/12, p. 1-9. CORREIA, A.A. Mendes (1938) – Raízes de Portugal. Lisboa : Edição da Revista Ocidente. CORREIA, V. (1921) - El Neolítico de Pavia. Madrid: Comisión de Investigaciones Paleontológicas y Prehistóricas, Memoria, 27. DIAS, Mª I; PRUDÊNCIO, Mª I; SANJURJO SÀNCHEZ, J; CARDOSO, G.O; FRANCO, D. (2008) – Datação por luminescência de sedimentos de sepulcros artificiais da necrópole pré-histórica da Sobreira de Cima (Vidigueira, Beja). Apontamentos

20 

 

no Alentejo Central: o contributo de Manuel Heleno. Lisboa: FLUL, master dissertation, 2 volumes, manuscript. ROCHA, L. (2007b) – O monumento megalítico do Lucas 6 (Hortinhas, Alandroal): um contributo para o estudo das arquitecturas megalíticas. Revista Portuguesa de Arqueologia. Lisboa: IPA, volume 10, nº 1, p. 73-94. ROCHA, L. (2009/2010) – As origens do megalitismo funerário alentejano. Revisitando Manuel Heleno. Promontoria. Faro: Universidade do Algarve, p. 4598. ROCHA, L. (2012) – Entre a vida e a morte: a perenidade dos espaços na Pré-História Recente no Alentejo (Portugal). Actas del 3 Simposium Internacional de Arte Rupestre. Cuba: Instituto Cubano de Antropología, XI Conferência Internacional Antropologia 2012. ROCHA, L; ALVIM, P. (2012) – Águias 2 (Brotas, Mora). Lisboa: IGESPAR, intervention report. ROCHA, L; ALVIM, P (2013) – Novas e velhas análises da arquitectura megalítica funerária: o caso da Mamoa do Monte dos Condes (Pavia, Mora). 5º Congresso do Neolítico Peninsular. Lisboa: UNIARQ, p. 521-527. ROCHA, L; FERNANDES, R. (2012) – Coast People, interior people... trade or breaks? In ALONSO RODRÍGUEZ, N.; ÁLVAREZ MARTÍNEZ, V; JIMÉNEZ CHAPARRO, J.I (Coord) – Actas del I Symposium Internacional “Gentes del Mar. Historia y Arqueología en el litoral del Arco Atlântico”. Pola de Siero, p. 327-331. SOARES, J. (2003) – Os hipogeus pré-históricos da Quinta do Anjo (Palmela) e as economias do simbólico. Setúbal: MAEDS. VALERA, A. (2008) – Recinto calcolítico dos Perdigões: Fossos e Fossas do Sector 1. Apontamentos de Arqueologia e Património. Lisboa: NIA-ERA Arqueologia, nº 3, p. 19-27. VALERA, A. (2009) – Estratégias de identificação e recursos geológicos: o anfibolito e a necrópole da Sobreira de Cima, Vidigueira. In BETTENCOURT, A. M. S.; BACELAR, L.; (eds) – Dos montes, das pedras e das águas. Formas de interacção com o espaço natural da pré-história à actualidade. Braga: CITCEM/APEQ, p. 25-36. VALERA, A; SOARES, A.M; COELHO, M. (2008) – Primeiras datas de radiocarbono para a necrópole de hipogeus da Sobreira de Cima (Vidigueira, Beja. Apontamentos de Arqueologia e Património. Lisboa: NIA-ERA Arqueologia, nº 2, p. 27-30. TRIGGER, B. (1992) – Historia del pensamiento arqueológico. Barcelona: Ed. Crítica VASCONCELLOS, José Leite (1929) – Antas da Herdade Grande. O Archeologo Português. Lisboa: Imprensa Nacional, volume XXVIII, p. 160-169.

21 

 

22 

 

Megalithic Rites of North Alentejo – Portugal Jorge de Oliveira Summary

The Micro-space

This article discusses the main evidence of the megalithic rites that have been recovered over various years of research in the area of North Alentejo, in Portugal. The big chronological gaps identified in dolmens and their connection to the Early Neolithic communities and to the menhirs (isolated or included in the dolmenic structures) are also analyzed. All of the absolute dates obtained up to the present for this region’s megalithic monuments are also presented.

Through an initial analysis, we can observe that the larger megalithic tombs apparently occur isolated. Nevertheless, if we prospect areas where the mechanization of agriculture has been less intense and, therefore, where the landscape is better preserved, in the territory immediately surrounding a dolmen we register other tombs of smaller sizes, usually less visible in the landscape. Although this “inter-visiblity” does not always occur, we can consider that isolated dolmens are very rare, and that they are usually organized in necropolises. But, at least in the area in question, these “cities of the dead” rarely have more than four tombs when located on class B and class C soils, and are usually built in granite. The necropolises located in worse agricultural quality soils (classes D and E), are generally schistose, can have a higher number of sepulchers, and one tomb of larger dimensions positioned at a more visible point of the landscape is always identified. It should be acknowledged that these observations result strictly from mere geographic analysis, and chrono-cultural or typological issues were not considered. Therefore, and for now, we are not taking into account if the monuments of one necropolis were all functioning at the same time, or if the filling up of one implied the construction of another.

Key-words: Megalithic radiocarbon dates.

culture,

rites,

Alentejo,

The macro-space Up to the present day, the architectural expression of death during the Neolithic in the North Alentejo area (Portugal) is almost exclusively identified in the common dolmens and their distinctive variations, which can have different dimensions, regular or elongated chambers, short or long corridors, and granite or schist as their main construction material. Six hundred and fifty of these monuments, isolated or forming a necropolis are known in the district of Portalegre. To understand their presence or their absence in a particular macro-area of this region, we have carried out various evaluations and published the results in different articles. Naturally, since the spaces for the deceased are near the spaces for the living, their implantation would be directly related to the existence of economic resources. These communities were the first to start depending on agriculture and pastoralism, so it is natural that they would settle on the lands most suitable for agro-livestock development, although in their own specific technological perspective. They thus looked for light and well drained soil, near waterlines of permanent course, where they could develop an agriculture that would be more like horticulture by today’s standards.

The positioning of the tombs on a macro-geographical scale having been clarified, it is now important to understand the reason or reasons for one monument to have been built in a certain place and not a few metres away. The theme of necessity and chance was broadly discussed in the 70s by Jacques Monot and it seems that it has been established that “chance” is something that does not exist in biological, let alone in psycosociological terms. Therefore, it is certain that one or more reasons, to a greater or lesser extent weighed up and justified by pragmatic issues, or broadly supported by mythological designs, would have existed for a tomb to be built in a specific location.

More suitable kinds of soil for pastoralism were also sought. The soils that are nowadays considered to be of high agricultural interest were clearly rejected. Despite being near waterlines, heavy and usually clayey soils were passed over by the Neolithic communities because they did not possess the technology that would allow them to work these lands. We have thus acknowledged that the soils nowadays classified as classes A and B were explicitly rejected by these communities. Having understood the strategy of the generic positioning of dolmens and correspondent settlements in terms of macro-scale, it is important to attempt to analyze if it is still possible today to identify some ritual standards for the micro-location of the deceased’s spaces and other ritual grammars.

Answers to this issue have already been occasionally tried out without significant results, being mainly based on astral or merely landscape-related observations. Studies carried out to approach this theme based on archaeological evidences observed in the paleo soils protected by the stone structures of the megalithic monuments’ mounds, at the bottom of the respective funerary spaces, or in the re-use of the tombs’ elements, are rare. In this article, based on our studies of megalithic funerary monuments carried out over a period of more than thirty years, we will try to gather information that might help us better understand some of the possible rituals, such as the ones that might be at the origin of the tombs, whose location did not result from a work of chance. 23 

 

Figure 3.1: Map of the Megalithic culture of North Alentejo.

Figure 3.2: Meada menhir (Castelo de Vide).

24 

 

all recognize how rare it is to find datable bone remains in schistose or granitic soils, soils which, due to their acidity, completely destroy organic material. We have most likely and successively been dating bones related to the final moments of the tombs’ use. It has to be kept in mind that the earlier dates, which are considered abnormal, invariably result from bits of charcoal that were systematically collected – when well preserved – at the bottom of the monuments or underneath the respective mounds. Because they went against the accepted theories, these dates were always rejected, being considered as related to pre-megalithic episodes and, therefore, much earlier than the construction of the tombs. The absence of studies of the earlier Neolithic habitats where equally old dolmens are located, and the fact that only recently have absolute dates of Early Neolithic contexts been made available, have contributed to the lack of interpretative proposals on a direct connection between dolmens, menhirs and the habitats of the first agro-pastoral communities. The studies we have developed in the Coudelaria de Alter area seem to make this connection easier, although, for now, it is based on a restricted group of elements. We shall now look into the parallels that it might be possible to establish between the materials collected at the Anta da Horta, located 500 meters away from the Porta do Tempo habitat. Inside the funerary chamber, where several re-visitations occurred, nine fragments with incised and imprinted decorations as well as plastic applications were identified; whose decorative motifs have direct parallels with the ceramics from the Toca da Raposa Locus and Locus II of the Porta do Tempo habitat. If parallels could be identified in the Reguengo Habitat’s ceramics, located fifty meters from this dolmen, we might be able to suggest that crosscontamination had occurred. However, the significant distance between these two places renders this hypothesis impossible. Amongst other examples, the resemblances between the ceramics from the Anta da Horta and from the Toca da Raposa are obvious and expressive. The same similarities can be seen in a perforated and subcircular section axe recovered from the Anta da Horta and in an axe with identical features, from the Toca da Raposa. But the most significant examples were identified in the Porta do Tempo’s Locus II. At the bottom of this small shelter, where decorated ceramics, manufacturing refuse and microblade technology were identified, a green rock pendant was recovered. In the more superficial layers of this shelter, in dragged earth, a fragment of what looks like an “ídolo-placa” (slab figurine), made out of sandstone and in its final manufacturing phase, was identified. From the examples given, it seems there are too many similarities between the materials of the Anta da Horta’s chamber and the obviously Early Neolithic materials of the Porta do Tempo habitat for us not to establish close connections between the users of the habitat and the builders or the first people buried in this dolmen. If for this dolmen we can identify, through the groups of artifacts, obvious connections with evidence attributed to very early moments of the Neolithic, in the Anta da Soalheira, in the Várzea Grande and in number 2 of the Vale de Carreiras, located in the Coudelaria de Alter, so phallic menhirs are

The times For a very long time we have passively accepted that dolmens and menhirs are contemporary. Nowadays, and especially after the dating of the Meada menhir (Castelo de Vide, Portugal), we know that menhirs are substantially older than the appearance of the dolmens. However, although that anteriority is today generally accepted, the identification of original or reused menhirs inside the dolmens or in their vicinities is growing in number. They were used either as structural elements or as apparently non-functional elements and therefore symbolic. In the region in question – North Alentejo – more than thirty menhirs, mainly isolated, have been identified. Amongst them, at least six are directly related to dolmens. The uncommonly early date of the Meada menhir, calibrated to a 2-sigma range between 5010 BC and 4810 BC, is, at first approach, clearly far from the 4th and 3rd millennium dates that usually frame the burials in megalithic tombs. If we accept that anteriority, the presence of menhirs inside the dolmens (which nowadays is starting to be recurrent) could have at least three possible explanations. The most simplistic one states that due to a resourcebased economy, the dolmens’ builders resorted to those already carved stones and used them for the construction of the tombs. A second hypothesis tells us that the inclusion of the menhirs might have been related to some ritual that would have obliged the dolmens’ builders to remove the menhirs and include them in their structure. A third hypothesis suggests that the dolmens were built in the place where a menhir already existed and it would be included in the structure, retaining the space’s sacralization. These hypotheses can probably coexist. Nonetheless, and regardless of the existence or absence of menhirs in the dolmens’ building structure, a series of dates, considered too early, have been available for a few years. These were mainly collected at the bottom of the dolmens in North Alentejo and in the Spanish “Extremadura”, and are close to the date of the Meada menhir and, simultaneously, to the dates generally available for the Early Neolithic habitats, which in calendar years refer to the 5th and sometimes to the beginning of the 6th millennium BC. Many explanations have already been presented in order to minimize the impact that these early dates would have on traditional interpretations of the origin of the megalithic culture, which always chronologically relates to a Middle and, more frequently, Late Neolithic. In fact, most of the available dates for dolmens are located between the 4th and the 3rd millenniums (in calendar years), although these dates were also mainly obtained from osteological material. However, nowadays it is fully accepted that these tombs had a very long working and functional life, with obvious signs of reuse, rehabilitations and developments. Despite this having rarely been done, it should be questioned whether the dates obtained from bone remains were not in fact related to the final moments of use and not to the foundational moment. We 25

 

found in the corresponding tomb structures, which also relate to early Neolithic phases. Given this situation, we should ask the question: when were the megalithic tombs of the Coudelaria Alter built? Certainly after the erection of the menhirs. But if we find mainly ceramic artifacts inside the dolmens attributed to the Early Neolithic, where should we position the menhirs’ phase? This “menhiric” phase cannot not be much earlier than the appearance of dolmens because the dates obtained from charcoal, collected at the bottom of the dolmens and underneath their mounds, such as in Castelhanas, Cabeçuda and Figueira Brava, in Marvão, and also in Joaninha, in Cedillo, put them in a time frame already available for the Early Neolithic habitats, the same occurring with the Meada menhir. Based on this data, we should accept that the “fashion” for the erection of menhirs was relatively short-lived, and quickly absorbed by the megalithic tombs. The menhir erection phase must have been very ephemeral or else the explanation for such early dates in the dolmens and the presence of ceramics of the Early Neolithic inside them has another explanation. The thesis on transference defended by Leonor Rocha (Rocha, 2005), is surely accepted in a broad sense, although it does not answer the question of the early dates obtained under the mounds. Prior to this, and based on obvious evidences, we have demonstrated that some of the dolmens studied in Northeast Alentejo were built on top of older habitats, where hearths and

silos occurred (Oliveira, 1997 and 1998). In other cases, however, this situation was not detected, the habitats being well defined, in the vicinity of the dolmens. The biggest problem relates to the relative positioning for the appearance of the menhirs. If, in fact, they existed before the building of the first dolmens, then during the Early Neolithic we should find, at least, two phases. An older one, during which the menhirs were erected and, afterwards, a phase of building megalithic tombs incorporating menhirs in the funerary structure. The alternative interpretation might one day be considered if the term “Middle Neolithic” is accurately defined. To fully re-evaluate this issue, the area of excavation of the habitats must be expanded and, above all, datable material should be sought to help us clarify with precision the periodization of the Neolithic in inner Alentejo. Through the available radiocarbon dates (although there are only two for the tombs in the schistose area), we have verified that these fall into the average chronological values for granitic monuments. Two charcoal samples collected in the Joaniña dolmen, located in the Termo Municipal of Cedillo, provided the following ages, respectively: sample A – 3840 ± 170 years BP; sample B – 5400 ± 210 years BP. The first sample relates to charcoal identified above the paved base of the monument, associated with an amphibolite axe and to an

Figure 3.3: Merliça dolmen (Castelo de Vide). 26

 

MAIN DATES FOR THE MEGALITHIC MONUMENTS OF THE NORTH ALENTEJO (PORTUGAL) MONUMENT 2 SIGMA COMMENTARY Laboratory AGE BP cal BC Anta da Horta – Human bones in the chamber associated Coudelaria de Alter 4190 ± 50 2930 – 2860 with a schist plaque. Undetermined (sample AH – M11) corridor dolmen. BETA - 194312 Anta da Horta – Human bones in the corridor’s entrance Coudelaria de Alter 4390 ± 40 3350 – 3020 associated with schist plaques, (sample AH-O10) sandstone, axes, keeled ceramics, BETA - 194313 bowls, etc. Anta das Castelhanas – Charcoal at the bottom of the chamber, Marvão 6360 ± 110 5448 – 5059 associated with straight or convex base (sample 2) arrow heads, made of schist. Short ICEN - 1264 corridor dolmen. Anta das Castelhanas – Carbonized human bones associated Marvão 3220 ± 65 1630 – 1380 with smooth semi-spherical pots, flint (sample I) 1340 – 1320 convex base arrow heads and a OXA – 5432 fragment of a schist plaque without decoration. Short corridor dolmen. Anta da Bola da Cera – Carbonized human bones associated Marvão 4360 ± 50 3258 – 2900 with an anthropomorphic shaped schist ICEN – 66 plaque, at the bottom of the chamber. Short corridor dolmen. Anta da Cabeçuda – Charcoal inside the “silo” at the bottom Marvão 3650 ± 110 2328 – 1736 of the chamber, associated with smooth (sample 1) 1715 – 1698 surfaced open bowls. Short corridor ICEN – 977 dolmen. Anta da Cabeçuda – Charcoal inside the chamber above the Marvão 7660 ± 60 6593 – 6577 granitic soil. Short corridor dolmen. (sample 2) 6564 – 6378 ICEN – 978 Anta da Cabeçuda – Carbonized acorn under the fallen Marvão 3720 ± 45 2274 – 2252 upright supporting stones, associated (sample 3) 2204 – 1971 with fragments of smooth surfaced ICEN – 979 ceramic. Short corridor dolmen. Anta dos Coureleiros – Charcoal inside the corridor associated Castelo de Vide 4240 ± 150 3335 – 2459 with a schist plaque with geometric ICEN - 976 shape and decoration. Long corridor dolmen. Anta da Figueira Charcoal at the bottom of the mound Branca – Marvão 6210 ± 50 5302 – 5270 derived from a non structured hearth, ICEN – 823 5014 – 5007 associated with a fractured lower grinding stone. Short corridor dolmen. Anta da Joaninha – Charcoal above the paved base of the Cedillo 3840 ± 170 2870 – 2800 monument, associated with an (sample A) 2770 – 2720 amphibolite axe, an arrow head and a Sac – 1381 flint blade. Schist dolmen. Anta da Joaninha – Charcoal between the monument’s Cedillo 5400 ± 210 4710 – 3770 paved base and the granitic soil, in (sample B) compact earth without archaeological Sac – 1380 material. Menir da Meada – Charcoal at the bottom of the Menhir of Castelo de Vide 6022 ± 40 5010 – 4810 Meada, inside the monument’s UtC – 4452 settlement cavity.

27 

 

ridges and did not include clear lithic features. The small size of these tombs contrasts with their density in each necropolis. Occasionally, more than ten of these tombs occupy the ridge of a small elevation. These structures, generally visible from each other, sometimes crown the top of all the hills of a small stream’s hydrographic basin.

arrow head and flint blade. The second sample relates to charcoal located between the paved base of the monument and the schistose soil, in very compact clayey earth, without any associated archaeological material. If sample B seems to correspond to charcoal prior to the monument’s construction, the second one is perfectly situated in regional megalithic contexts, very close to the available date for the Anta da Bola Cera (a granitic monument with a short corridor), located in the municipality of Marvão. This sample, obtained through burnt human bones, associated with an anthropomorphic shaped schist plaque, provided the following age: 4360 ± 50 years BP. Likewise, sample 1 from Anta da Cabeçuda, a monument with a short corridor, located in the municipality of Marvão, refers to charcoal collected at the bottom of the chamber, associated with open bowls, that provided the following age: 3650 ± 110 years BP. Another monument, this time with a long corridor – the Anta IV dos Coureleiros – located in the Castelo de Vide municipality, provided a sample of charcoal collected from the corridor, associated with a geometric shapedschist plaque, with the following date: 4240 ± 150 years BP. Therefore, a perfect framework for sample A from Anta da Joaniña is recognized in the same chronological contexts as the granitic megalithic monuments of North Alentejo. Sample B from Anta da Joaniña is equally framed within the group of so-called old dates of the granitic area’s megalithic culture. In this group, the following values are known: Sample 2 from Anta das Castelhanas: 6300 ± 110 years BP; Sample 2 from Anta da Cabeçuda: 7660 ± 60 years BP; and Anta da Figueira Branca: 6210 ± 50 years BP.

The effort necessary for the construction of the schist tombs involves a tiny fraction of the effort spent on one of the granitic tombs. The time, the strength and consequently the prestige necessary to gather them together would have been significantly different in each community. However, the architectural discourse present in both areas under analysis follows similar patterns; although clearly poorer in the case of the schist constructions, the funerary equipment holds the same message; its state of conservation is the same as that of the secondary deposits identified in the granitic area. Therefore, it seems that in generic terms and in the face of the deficient data we possess, the ritual environment identified in all the megalithic culture of North Alentejo follows the same patterns, although each region expresses them in its own specific way. Although the essence of the discourse seems to be the same, we find, even within each group, various ways of addressing death, it thus being possible to isolate different types of rituals. There have been identified, regardless of the monuments’ architecture or their spatial positioning, different forms of entombment. In the granitic region, we find testimonies of various types of dolmens – dolmens with primary deposits close to the ground, on which pouches with human remains prepared outside were deposited; dolmens which seem to have always served exclusively as ossuaries; dolmens whose interiors, as well as serving as ossuaries, were also used for the preparation of corpses; and dolmens that continued to receive entombments in much later periods.

It seems, therefore, that there are no doubts and that based solely on two dates from Anta da Joaniña, the small megalithic monuments from the Sever river mouth are the contemporaries of larger monuments, located on the granitic plane of the S. Mamede mountain range. In the light of the available data, it seems there are no doubts regarding the contemporaneity of both groups of monuments. The dissimilarities between the architecturalvolumes and the sets of artifacts would result from the different socio-economic organizations imposed by the specificity of existing resources in each area.

The disposition of primary as well as secondary funerary deposits inside the funerary space seems to have followed pre-established norms. The placing in a privileged position, generally near the chamber’s entrance, of at least one plaque made out of mica schist or sandstone, with an anthropomorphic outline or decoration, which stands out from the others due to its finishing touch, has been identified too many times to be considered accidental. This placing has been confirmed in granitic monuments, as well as in the Pombais and Fonte da Pipa dolmens, located in the schist area. The stones that encompassed these big plaques seem to have functioned as small altars or niches to make them stand out. On these small stages, the remains of painting, or a discrete dusting of ochre or (in rarer cases) cinnabar, are always present.

The modes The small number of excavations carried out on schist monuments and the limited information that the dolmens studied have provided, render the complete identification of funerary practices performed in these monuments very difficult. Taking into account the funerary offerings and also the human remains identified in either of the two large versions of tomb, in schist or in granite, they are collective funerary spaces. In the course of our research, we have realized that exhumed materials from schist tombs are, clearly, poor when compared to those collected from larger monuments located in the granitic region. The poorness of the spoil is also reflected in the volume of the tomb, which in most cases would pass unnoticed in the landscape if it did not occupy mountain

Although in some cases the effects of violations might have contributed to a dispersal of the materials to the chambers’ periphery, in other monuments where 28 

 

fractured, the slate schist plaques, of mica schist, or sandstone, are always found at the monuments’ bases. Evidences of the plaques’ reuse in funerary monuments are also well documented in this region. Seven plaques were recovered from the Tapadão da Relva dolmen, excavated by the former Grupo de Arqueologia de Castelo de Vide (Castelo de Vide’s Archaeology Group), two of which showed clear signs of reuse. TR 24 constitutes an interesting example of the re-use of a larger plaque, which was partially recovered, certainly after fracture, in order to obtain a new plaque. Of the original piece it is still possible to see somewhat erased zigzag stripes, which in the new plaque form an obviously unbalanced result. A new hole with bifacial perforation was made on the top of the plaque by the people who reused it, and they simply regularized its shape, which further contributed to some of the engravings’ disappearance. TR 52 might be one of the best examples of the reuse of an accidentally fractured plaque. The larger part of this plaque survived. On the bottom, near the fracture line, two small bifacial holes can be seen near the lateral contours. These were surely made in an attempt to put back together the two pieces into which the plaque had broken. Unfortunately, only one of the fragments was recovered.

violations have not been registered, we still continue to identify that the largest concentration of both bones and spoils mainly occurs near the upright stones that support the structure. In some cases (Horta, Anta dos Pombais, Bola da Cera, S. Gens II, Padre Santo, Castelhanas, etc.), various votive materials were clearly positioned between the blocks steadying the upright stones. The pots that were positioned vertically between the blocks are of particular importance, since they contrast with the inverted position of those found elsewhere. It has thus been noted that both the funerary pouches and the primary inhumations were preferably carried out near the chambers’ upright stones. In the rare cases where it has been possible to understand the anatomical relation of the osteological elements of the primary tombs, we have verified that the long bones were usually located nearer to the upright stones than the skulls. This finding might not only be explained by the corpses’ disposition in fetal position and lateral decubitus, as occurred in the Bola da Cera dolmen, but also by the corpses, initially leaning against the upright stones, falling over. Assuming that the slab figurines were mainly placed on the necks of the buried corpses, this would explain the face down position in which most of the plaques were found.

In Anta I dos Coureleiros, excavated by us in 1991, a schist plaque was found. Through its still preserved odd jagged partial outline, it seems this plaque was manufactured from a fragment from a crosier and it is probably similar to the one collected at Anta Grande da Herdade das Antas, in the Montemor-o-Novo municipality.

The information recovered from the monuments’ corridors, is smaller in quantity and less explicit. In none of the cases were we able to reliably identify primary deposits in the corridors. When present, the bone remains either contained fire evidence (as in Castelhanas) or were limited to small splinters. Although in some corridors the recovered materials were well preserved (such as the one from the Coureleiros IV’s corridor) nothing proves that they accompanied primary deposits. At this point, the presence of a big, low keeled vase, which lacks a fragment, and which points to Late Neolithic or even Chalcolithic chronologies, might indicate late or probably secondary tombs. However, the date obtained for the charcoal associated with a schist plaque from the same level as this vase and recovered from less than 150 meters was 4240 ± 1 50 years BP – too early to confirm the previous observations. This diversity of situations might be explained by the constant visitations that these monuments witnessed and that were not always identifiable through the earth’s consistency or coloration

Besides this plaque, another three whole ones and a fragment were recovered in Anta I dos Coureleiros. When compared to the other one, two of these plaques showed a slightly exaggerated curve on one of their edges, which suggest that they might also have been produced from the same or another crosier, as CI 4 certainly was. These plaques’ re-uses, “collages” and re-engravings clearly show that entombment was not seen as the end of the individual’s path. The plaques’ breaking and placing together again (well shown on TR 52), the re-engraving of some of the pieces from the Crato, Alter and Elvas municipalities, and the likely transformation of crosiers into new plaques (such as the ones collected from Coureleiros I), might constitute examples of how the tombs were in fact living spaces. The manipulation of bone remains in different phases is well documented in Anta da Horta, in Alter do Chão. A skull and a calcaneum were removed from the monument’s chamber, transported and organized at the corridor’s entrance, probably in an entrance hall. The calcaneum supported the skull and around them was a large number or sandstone plaques, with and without decoration, ceramic vessels, axes and adzes, arrow heads and a striated pin head fragment made of bone. This was clearly an act of tribute paid to a corpse buried earlier, carried out in a much later period after his death.

If we exclude the striking funerary deposit in a probable entrance of Anta da Horta, located in the Coudelaria de Alter, it is worth mentioning the absence of sandstone or mica schist plaques, anthropomorphic or not, in undisturbed levels of the monuments’ corridors. In the other cases where this kind of material was found (Anda da Cabeçuda, Pombais and Coureleiros II); they were located in earth that had been broken up and obviously dragged out from inside the chamber. On the other hand, the presence of schist plaques is documented in both chambers and corridors, although we ascertain that in monuments with well preserved levels, the schist plaques with exclusively geometric shape and decoration are mainly found in the corridors. When complete or slightly

29 

 

  Figure 3.4: ‘Ídolos-placa’ or slab figurines (Horta dolmen – Coudelaria de Alter).

Figure 3.5: Burnt human bones (Castelhanas dolmen – Marvão). 30 

 

Figure 3.6: Toasted acorn (Cabeçuda dolmen – Marvão). other monuments, together suggest that at least some megalithic tombs were built over older habitats. This continuing occupation of the same space, initially by the living and afterwards by the dead, occurs too frequently to happen by chance. The symbolism of a place, which is probably related to the appropriation of a territory by the group’s ancestors, continued by the superimposition of a tomb with all the symbolic weight that this naturally possesses and by continual visits, would transform this place into an archive where the group’s memory would be perpetuated. The abundant presence of elements of mills in the burial structures might be related to the presence of the earlier habitats. The intentional fracture of these essential elements of transformation and their inclusion in the building of the monument seem to surpass the simple act of adapting a stone for use as a steadying block for an upright stone. A deeper meaning certainly involves the fracture of the elements of mills placed in funerary spaces, but for now, this is hard to understand.

The frequent absence in funerary spaces of fragments of votive objects in sealed levels of secondary depositions indicates that the external preparation of corpses was already accompanied by these objects. If this explains the absence of a large number of fragments we can not, however, disregard the fact that during the various journeys that the funerary pouches certainly made, some of the already affected objects would have easily been lost. During the multiple visits that occurred during the monuments’ working life, it is also likely that some of the materials, or their fragments, might have been removed because of the symbolic weight they possessed. These constant visits are also attested in the materials (some of them very late) that were collected from the mounds. On the outside of the dolmens, axe deposits, fragments of chalcolithic plates, Bronze Age vases and even roman and some medieval ceramics were collected. These late materials were probably a result of offerings and testify the continuity of the symbolic weight in the funerary spaces. We should also keep in mind the Bronze Age funerary deposit inside the Bola da Cera dolmen’s chamber. If the funerary space projects itself well after the time when it was built and initially used, it seems to preferably establish itself in already humanized spaces. The dating of charcoal, surrounded by various ceramic fragments, collected inside a non structured hearth at the bottom of the mound of the Figueira Branca’s dolmen (in Marvão), the identification of the bottom of a hut at the base of Huerta de las Monjas’s mound (in Valência de Alcântara), the dating of charcoal collected at the base of the Cabeçuda dolmen’s chamber (in Marvão), and also the various reports of human occupation identified under

The presence of hyaline quartz crystals and above all of tourmaline crystals in most of the dolmens seems to be non accidental. Both in the schist and the granitic monuments (but mainly in the granitic ones), big quartz crystals, some of them tinted, usually occur near the lower levels. The tinted hyaline quartz crystals collected in Coureleiros III show signs of abrasion on the extremities. The signs of use identified in these two pieces are not present in the crystals collected in other monuments. Therefore, it seems that the inclusion of these geological rarities near the deceased could be understood as an ornament or ritual object and not as an object of daily use. 31 

 

Figure 3.7: Funerary offerings. The understanding of other less common manifestations identified in this region’s monuments is also a difficult task. We shall approach some of the cases. In the Pombais dolmen, at the lever where the earth is more compact, three small pebbles, beige in colour and eggshaped, were collected. Two of them showed signs of having been sprinkled with ochre or cinnabar. The eggshape of these small pebbles might have a similar meaning to the lagomorphs, of which only one is known in this region, found in the Tapada de Matos dolmen (in Castelo de Vida), but which are well represented in Central Alentejo and Estremadura. The longitudinal opening on adults’ long bones identified in the Castelhanas dolmen, the presence of roasted acorns, of a pebble and an axe collected at the bottom of a likely posthole, identified in the centre of the Cabeçuda dolmen’s chamber, the human foot-shaped block of granite identified in the area of the Coureleiros III mound, the circular structure registered in the Figueira Branca dolmen’s mound (in Marvão), the rectangular one detected inside the Bola da Cera’s chamber (Marvão), or the one identified in Charca Grande de la Regañada (Cedillo), the frequent use of ochre and other above mentioned aspects whose function is difficult to explain, confer a deeply symbolic meaning to the group of materials and structures that these monuments held and that, millions of years later, only gradually and with difficulty can be decoded.

Conclusion A large amount of information has yet to be fully understood, but some of the data seems to be already confirmed. It is today commonly accepted that the megalithic monuments with funerary characteristics only had the capacity to hold a small part of the community’s deceased. The indicated exclusivity of materials collected in the different monuments of the Coureleiros necropolis (studied by us) and the different volumes of its dolmens seem to suggest a social differentiation amongst the buried. Although we did not collect bone material from this necropolis, the material collected in other monuments shows us that the selection of those entitled to be entombed in these monuments took no account of either age or, very likely, gender. Despite the unreliable data, some of the bone remains identified in the Bola da Cera dolmen’s pouches might have belonged to female adults. We have verified that various bones from different individuals and different monuments showed signs of a life subjected to very violent episodes that, in some cases, lead to death. The trauma caused by cutting or heavy objects showed that at least some of the entombed had been clearly exposed to very violent situations. It was found that the higher percentage of traumatized bones belonged to adult individuals, but not elderly 32 

 

do concelho do Crato (Alto Alentejo) – V. Trabalhos de Antropologia e Etnologia. Porto: SPAE, volume 17. ISIDORO, A. F. (1975) – Escavações em dólmenes do concelho do Crato (Alto Alentejo) – VI. Trabalhos do Instituto de Antropologia Dr. Mendes Correia. Porto: Universidade do Porto, nº 29. LEISNER, G.; LEISNER, V. (1956) – Die Megalithgraber Iberischen Halbinsel. Der Westen. Berlin: Walther de Gruyter, nº 1. LEISNER, G.; LEISNER, V. (1959) – Die Megalithgraber Iberischen Halbinsel. Der Westen. Berlin: Walther de Gruyter, nº 2. LEISNER, G.; LEISNER, V. (1965) – Die Megalithgraber Iberiscishen Halbinsel. Der Westen. Berlin: Walther de Gruyter, nº 3. MONTEIRO, J. P.; GOMES, M. V. (1977) – Os Menires da Charneca do Vale do Sobral – Nisa. Revista de Guimarães. Guimarães: Sociedade Martins Sarmento, nº LXXXVII. OLIVEIRA, J. (1993a) – Territórios e Variabilidade Megalítica no Nordeste Alentejano. Actas do 1º Encontro Transformação e Mudança. Cascais/Lisboa: UNIARQ. OLIVEIRA, J. (1993b) – Reutilizações e Reaproveitamentos de Materiais em Sepulturas Megalíticas do Nordeste Alentejano. Actas do 1º Congresso de Arqueologia Peninsular. Porto: ADECAP, volume I. OLIVEIRA, J. (1995a) – A Recuperação do Menir da Meada – Castelo de Vide. Ibn Maruán. Marvão: Câmara Municipal de Marvão, nº 5. OLIVEIRA, J. (1995b) – Sepulturas Megalíticas del Termino Municipal de Cedillo - Província de Cáceres. Cáceres: Edicion del Ayuntamiento de Cedillo. OLIVEIRA, J. (1995c) – A Recuperação do Menir da Meada - Castelo de Vide. Castelo de Vide: Câmara Municipal de C. de Vide. OLIVEIRA, J. (1996a) – Datas absolutas de monumentos megalíticos da bacia hidrográfica do Rio Sever. Actas do 2º Congreso de Aqueologia Peninsular. Zamora. OLIVEIRA, J. (1996b) – As pequenas antas de Montalvão e Cedillo. Actas do I Colóquio Internacional sobre Megalitismo de Monsaraz. Lisboa: Câmara Municipal de Reguengos de Monsaraz/UNIARQ. OLIVEIRA, J. (1997) – Monumentos Megalíticos da Bacia Hidrográfica do Rio Sever. Lisboa: Colibri, 1º Volume, bilingue edition, sponsered by the councils of Marvão, Castelo de Vide, Nisa, Valencia de Alcântara, Herrera de Alcântara and Cedillo; also Delegação Regional do Ministério da Cultura. OLIVEIRA, J. (1998) – A Anta de la Joaniña e a da Era de los Guardias no ambiente megalítico da foz do Sever. Ibn Maruán. Marvão: Câmara Municipal de Marvão, nº 8. OLIVEIRA, J. (1999) – Inventario, Investigacion y puesta en Valor de los Dólmenes: Termino Municipal de Cedillo. Extremadura Restaurada.

individuals. These observations show us that some of the entombed might have been directly involved in warlike dealings, showing us that the communities of builders and users of the megalithic burials would not have lived in a completely pacific environment, as the absence of fortified settlements might suggest. Although we do not know to whom these tombs were dedicated, we do know that they were multi-functional. Some of those superimposed on habitat soils seem to guarantee the continuity of the territory’s occupation; prepared to receive primary burials, they also witnessed the preparation of corpses and complex funeral ceremonies, they functioned as ossuaries and, from the late materials, we know that they were continuously visited and transformed into likely meeting places. Mainly built with the strength of many people, they held only those that had enough prestige to warrant the eternity of a space. References BUENO, P. (1986) – Megalitos en Extremadura. Actas de la Mesa Redonda sobre Megalitismo Peninsular. España/Portugal, 1984. BUENO, P. (1987) – Megalitismo en Extremadura: Estado de la Cuestión. El Megalitismo en la Península Ibérica. Madrid: Ministerio de Cultura. BUENO, P. (1988) – Los Dolmenes de Valencia de Alcantara. Excavaciones Arqueologicas en España. Madrid: Ministerio de Cultura, nº155. BUENO, P. (1989) – Camaras Simples en Extremadura. XIX Congreso Nacional de Arqueologia. Castellón de la Plana, 1987. DIAS, A. C.; OLIVEIRA, J. M., (1981) – Monumentos Megalíticos do Concelho de Marvão. Portalegre: Assembléia Distrital de Portalegre. HENRIQUES, F. J. R.; CANINAS, J. C.; CHAMBINO, M. (1993) – Carta Arqueológica do Tejo Internacional. V.V.de Ródão: A.E.A.T, volume 3. ISIDORO, A. F. (1966) – Escavações em dólmenes do Concelho do Crato (Alto Alentejo). Trabalhos de Antropologia e Etnologia. Porto: SPAE. ISIDORO, A. F. (1966) – Contribuição para o Estudo da Arqueologia do Concelho de Alter do Chão (Alto Alentejo). IV Colóquio Portuense de Arqueologia. Guimarães: Sociedade Martins Sarmento, Revista de Guimarães, Porto, 1965. ISIDORO, A. F. (1967) – Escavações em dólmenes do Concelho do Crato (Alto Alentejo) II. Trabalhos de Antropologia e Etnologia. Porto: SPAE, volume 20. ISIDORO, A. F. (1970) – Escavações em dólmenes do Concelho do Crato (Alto Alentejo) III. Anais da Faculdade de Ciências. Porto: Universidade do Porto, nº 54. ISIDORO, A. F. (1971) – Escavações em dólmenes do Concelho do Crato (Alto Alentejo) IV. Trabalhos de Antropologia e Etnologia. Porto: SPAE, volume 22. ISIDORO, A. F. (1973) – Escavações em dólmenes 33 

 

Mérida: Consejería de Cultura y Patrimonio de la Junta de Extremadura. OLIVEIRA, J. (2000a) - Economia e sociedade dos construtores de megalitos da bacia do Sever. Actas do 3º Congresso de Arqueologia Peninsular. Neolitização e Megalitismo da Península Ibérica. Porto: ADECAP, Vila Real 1999, volume 3, p. 429-444. OLIVEIRA, J. (2000b) – Continuidade e Rupturas do Megalitismo do Distrito de Portalegre. Actas do 3º Congresso de Arqueologia Peninsular. Neolitização e Megalitismo da Península Ibérica. Porto: ADECAP, Vila Real 1999, volume 3, p. 459-471. OLIVEIRA, J. (2000c) – A Anta II de S. Gens – Nisa. Ibn Maruán. Marvão: Câmara Municipal de Marvão, n.º 9/10. OLIVEIRA, J. (2000d) – A Anta da Tapada de Matos – Castelo de Vide. Ibn Maruán. Marvão: Câmara Municipal de Marvão, n.º 9/10. OLIVEIRA, J. (2001) – O Megalitismo de Xisto da Bacia do Sever – Montalvão Cedillo. Muitas antas pouca gente? Trabalhos de Arqueologia. Lisboa: Instituto Português de Arqueologia, Actas do I Colóquio Internacional sobre Megalitismo, 16, p. 135-158. OLIVEIRA, J. (2004) – O Megalitismo do Distrito de Portalegre 100 anos depois do inventário de Francisco Tavares de Proença Júnior. Arqueologia: Colecções de Francisco Tavares de Proença Júnior. Lisboa: Instituto Português de Museus. OLIVEIRA, J. (2006) – Património Arqueológico da Coudelaria de Alter. Lisboa: Colibri/Universidade de Évora. OLIVEIRA, J. (in press) – Coudelaria de Alter – 3 anos de trabalhos arqueológicos. Actas das 3as. Jornadas de Arqueologia do Norte-Alentejano, 2007. OLIVEIRA, Jorge de (2008) – The Tombs of the Neolithic Artist-Shepherds of the Tagus Valley. In BUENO-RAMIREZ, P.; BARROSO-BERMEJO, R.; BALBÍN-BERHMANN, R. (coord.) – Graphical markers and megalith Builders in the International Tagus, Iberian Peninsula. Oxford: BAR Publishing, BAR International Series 1765. OLIVEIRA, J. (2012) – Monumentos Megalíticos da Bacia Hidrográfica do Rio Sever. Évora/Marvão: Chaia/Câmara Municipal de Marvão, 2nd and 3th volumes, bilingue and electronic edition). OLIVEIRA, J.; DIAS, A. C. (1981) – Monumentos Megalíticos do Concelho de Marvão. Portalegre: Edição da Assembleia Distrital de Portalegre. OLIVEIRA, J.; MOITAS, E; OLIVEIRA, C. (in press) – Monumentos Megalíticos do Concelho de Arronches. Actas das 3as. Jornadas de Arqueologia do Norte-Alentejano, 2007. OLIVEIRA, J.; OLIVEIRA, C. (2000) – Menires do Distrito de Portalegre. Extremadura Arqueológica. Cáceres: Universidad de Extremadura, Número Especial de Homenagem a Elias Diegués. OLIVEIRA, J.; PARREIRA, J.; PEREIRA, S. (2007) – Nova Carta Arqueológica de Marvão. Ibn Maruán. Marvão/Lisboa: Câmara Municipal de Marvão/Colibri, edição especial.

OLIVEIRA, J.; RIBEIRO, M.; PINTO, M. (2007) – Património Arqueológico em Nisa – Revisão do PDM. Actas das 3as. Jornadas de Arqueologia do Norte-Alentejano. RODRIGUES, M. C. M. (1975) – Carta Arqueológica do Concelho de Castelo de Vide. Lisboa: Assembleia Distrital de Portalegre. English language version: Sofia Lovegrove ([email protected])

34 

Death as “life’s mirror”? Funerary practices and trajectories of complexity in the prehistory of peasant societies of Iberia João Carlos SENNA-MARTINEZ

Abstract

Resumo

In Iberia, between Middle Neolithic and the Late Bronze Age, funerary practices will suffer transformations that encompass the main structural change of society. I.e. such changes sometimes allow and can even substantiate periodization options. In this paper we propose to discuss such dynamics, their archeographic visibility and consequences to the construction of social models.

Na Península Ibérica, entre o Neolítico Médio e o Bronze Final, o tratamento da morte sofre transformações que acompanham as grandes rupturas estruturais das sociedades. Isto é, permitindo e substanciando, por vezes, opções de periodização. Propomos uma reflexão sobre estas dinâmicas, a sua visibilidade arqueográfica e respectivas consequências para uma modelização social.

Key-Words: Western Iberia, Prehistory of Peasant Societies, Funerary Practices, Social Models.

Palavras-Chave: Ocidente Peninsular, Pré-História das Sociedades Camponesas, Praticas Funerárias, Modelos Sociais.

Besides iconography, the main window that allow us to peek into social change and complexity transformations in Iberia during the Prehistory of Peasant Societies (i.e. the historical phase that encompasses the period of time between Early Neolithic and the Late Bronze Age) is funerary practice. This prehistoric phase can be divided into two longduration trends: (1) the one we can generically call “Neolithic” and which we perceive as encompassing part of the so called Chalcolithic period; (2) the Bronze Age with some aspects of the Late Chalcolithic.

Funerary practices of Iberian Early Neolithic societies are not yet very well known, namely in Portugal. If we take as representative the case of Castelo Belinho’s Neolithic village in Western Algarve (Portugal) it shows an estimated population of “…between thirty and fifty people…” and fourteen documented graves, interspersed between the eight habitational structures excavated (Gomes, 2012, p.113-115) with a presumed social organization that possibly “…relied on a communitarian model based on several household units…” (Id.). In the Spanish Mesetas a few recent finds show a similar funerary situation in the Early Neolithic with scarce and isolated graves, mostly in pits, always in domestic open air habitats or within caves (Rojo-Guerra and GarridoPena, 2012).

1. In the Beginning If we accept Gimbutas proposal of the role of a “Goddess” in the origin of a “matrifocal and probably matrilineal, agricultural and sedentary, egalitarian and peaceful” society (Gimbutas, 1996, p.9) then we must think how this general model can or can’t be adapted to the actual conditions of Iberia regions in the times and historical phasing of pre-urban peasant societies.

Similar burials are known at the Moura and Serpa areas in Alentejo, but at later dates. Most of the pit burials occur in negative structures of inverted bell shape like grain silos usually are. This situation probably implies re-use of previously existing silos but can also have a link with the metaphor for life regeneration that the concept of seed storage implies (Williams 2003, 242).

The Early Neolithic in Iberia is not uniform across its different regions. As it will happen during the lapse of time encompassing the existence of its prehistoric peasant societies (from Early Neolithic to the Late Bronze Age) Iberian regional diversity will be the rule. Namely, the extent of the Neolithic agricultural component will vary profoundly from area to area in Iberia, with cereals being the latecomers for its central and northwestern areas (Zapata, et al. 2004).

Whenever their conservation state allows a complete contextual analysis, this type of burial implies structures that were closed thereafter, i.e. which were not opened again to dispose new individuals.

Even in the Levantine area (the shores of Cataluña and Valencia) it seems that the early agro-pastoralists to arrive coexisted for a time with “indigenous” mobile hunter-gatherers, till these last ones are “converted” to the new ways (Id.; Fairén-Jiménez, 2007).

In Castelo Belinho, the possibility of a spatial layout of the funerary pits and houses according to probable “astronomic orientations” (Gomes, 2012, p.115), eventually of a seasonal calendar nature, opens the door 35 

 

either to land cultivation, pastures and or foraging rights, alone or all together, so being capable of a multiplicity of socioeconomic combinations encompassing a large spectrum of resources that have in common a new territorial approach. Several areas in Iberia see this happen between the last quarter of the 5th millennium BC and the first quarter of the 4th. In Atlantic Iberia – Central and North Portugal and Galicia – as well as the interior plateaus the season of critical resource availability would clearly be winter. Then lowland pastures and fall foraging would support both flocks and men and complement whatever contribution a small-scale agriculture could provide. In Beira-Alta plateau acorn recollection was proved to be such an important winter resource whence toasted and stored (Senna-Martinez, 1995; Senna-Martinez and Ventura, 2008a and 2008b).

to a possible correlation with what happens with the development of megalithic monuments from the Middle Neolithic onwards. 2. From Middle Neolithic to the Chalcolithic: In between Megalithism. Since the transition to farming and pastoralism occurs at a different pace and in multiple forms, in the different regions of Iberia, combined or not with hunter-gatherer practices, it’s by no means a sure thing that agriculture is everywhere in Iberia a pre-condition to the emergence of the construction of megalithic funerary monuments. In several areas it can be argued that animal husbandry predominates (namely ovi-caprids) together with different degrees of what we can call horticulture, hunting and gathering. Such is the case of Central Portugal (SennaMartinez and Ventura, 2008b).

Fall acorn and other winter fruits gathering and processing seems to go together, in several Iberian regional areas, with the period (fall and winter) when building of megalithic tombs took place, either we agree with the solar hypothesis for their orientations (Hoskin, 1998, 2001) or take into consideration other proposed astronomical orientations (Silva, 2012).

What the different regional scales of agriculture and animal husbandry adoption have in common with specialized gathering (such as acorns as in Central Portugal – Id.) is the development of a proprietorial attitude toward land (Fairbairn, 2000), leading to what we have called the passage from “use-territories” to “occupation territories” (Senna-Martinez and Ventura, 1999).

In the Mondego’s river interior basin (or plateau) fall and winter seem to be the seasons when lowland habitats were in use (Senna-Martinez and Ventura, 2008b) bringing into being the regional seasonal pattern of transhumance between lowland and highland pastures that persisted into contemporary times (Martinho, 1981). The return to earth and rebirth metaphors for burial that go together with the agricultural cycle as metaphor (Williams, 2003) can be used to explain the birth of megalithic tombs as a natural cave replacement in burial, and also explain why in many areas of Iberia where natural caves exist these were in such use for the Prehistory of Peasant Societies duration and sometimes longer. For instance, ritual similarities between megalithic burial and cave burial in Portuguese Estremadura leaded Victor Gonçalves (1978) to support the concept of “cave megalithism” to emphasize such similarities (Boaventura, 2009). Late Neolithic and Chalcolithic will bring both some trend confirmation and transformations to megalithism and burial practices.

 

Figure. 4.1: The Middle Neolithic Dolmen of Folhadal built over the hut remains of an Early Neolithic habitat. We defended this to be the probable main reason for the emergence of Megaliths (fig.1), which marks the transition between Early and Middle Neolithic in the Mondego’s Platform in Central Portugal (Senna-Martinez and Ventura, 2008a), and we think that reasoning can be extended to the other Iberian areas where early funerary megaliths evolved.

The first transformation that occurs is in terms of scale. While the small monuments of Middle Neolithic seem to imply that not all the elements of the community that built them would be buried there (Senna-Martinez and Ventura, 2008b, p.325) the large passage-grave tombs of Late Neolithic (fig.2), now built in greater numbers, could accommodate a much larger number of burials accompanying the Neolithic “demographic revolution”. These later monuments also imply a larger community for their construction with more time and more hands needed. The calculations made for the Beira-Alta dolmens show a multiplication by more than three of the required time and work-force between Middle and Late

Critical resources imbalance and social dispute can be said to be behind the development of formal disposal areas for the dead (Chapman, 1981, p.80) as a form of legitimizing territory ownership. Ad monumentality to this equation and funerary monuments can be related 36 

 

Neolithic monuments (Senna-Martinez and Ventura, 2008b, p.337).

revolution surely reflected itself in the development of regional Chalcolithic settlement systems of increasing complexity that the last two decades have shown to encompass the treatment of the dead (Valera, 2012).

 

Figure 4.2: The Late Neolithic built Dolmen of Fiais da Telha. A large and well preserved megalithic tomb of the Mondego’s platform. The second generalized transformation relates to the artefactual “package” thought as adequate to follow the dead into the grave. From the small number of artefactual types present in the Middle Neolithic tombs to a situation where virtually all objects present in life could be found inside the tombs. Pottery is an important part of this “funerary package” transformation, being very scarce or even totally absent from the early megalithic burials it becomes abundant in the Late Neolithic and seems to replicate in death the “serving set” in use in everyday life, as was proved in the case of Beira-Alta communities (Fig.3 – Senna-Martinez and Ventura, 2008b, p.342).

Figure 4.3: Pottery types from the huts of Ameal-VI habitat in the Mondego’s Platform. The same types occur in the nearby megalithic tomb of Fiais da Telha.

We argued (Id., Ibid.) that the almost total absence of mobile female figurines and or representations, as a Late Neolithic and Chalcolithic funerary item, north of the Central Iberian Massif could be linked to different regional economic roles of women. We think that the full agricultural societies from the Levantine and Southern half areas of Iberia granted a correlative greater importance to women as an important part of the labour force involved.

In the Iberian Southwest,1 besides large dolmens and tholoi, different types of rock-cut tombs were built in areas were the bedrock allowed, as well as continuing the practice of pit and ditch burials that seem to originate in the Early Neolithic. This enlarged burial capacity, inside and around settlements, plainly justifies the concept of a “democratization of death” coined to characterize this demographic aspect of Chalcolithic burial practices. Furthermore, the recent discovery that to the already know walled Chalcolithic settlements we should add a large number of ditch-limited enclosures of different types, built and used during the Late Neolithic and Chalcolithic, profoundly changed our knowledge.

Elsewhere, as for example in the Portuguese Beira-Alta, more mobile economies based in animal husbandry, hunting and plant gathering, with small agricultural contributions, would lead to a lesser iconographic visibility of women’s role. Nevertheless, the important role of pottery as a funerary item from the Late Neolithic to Chalcolithic times, iconographic invisibility notwithstanding, can imply a growing feminine social role in this period.

Within some of these ditch-limited enclosures, as well as in the larger funerary monuments, we have proof that different kinds of treatment affected human remains both with primary and secondary depositions, sometimes with segmentation and rearrangement of bodies and even cremation (Valera, 2012, p.111; Silva, et al. 2012).

Secondary products revolution (Sherrat, 1981), and namely the invention of the plough, has been presumed to be one of the causal phenomena behind the development of full agricultural systems in Southern Iberia with extensive cereal dry-farming (Gonçalves, 1989). Probably occurring during the Late Neolithic, this socio-economic

                                                             1

This is a cultural area that encompasses the Portuguese Alentejo and Algarve Regions, as well as the western part of Spanish Andalucí a.

37 

 

 

During the Late Neolithic and the Chalcolithic, the Iberian SW is also the area where iconographic expressions of a feminine symbolic entity, that we can think of in terms of Gimbuta’s (1996) “goddess”, largely occur together with other expressions of “the agricultural cycle as metaphor”. In between cromlechs, megalithic chamber tombs astronomical orientations, and astronomical alignments of structural components in some ditch-limited enclosures, most of the symbolic components associated with life and death in these communities can be viewed as components of the symbolic of the agricultural cycle as metaphor we already discussed. These symbolic expressions probably reflect a long trend world view, we can call Neolithic, that we think characterize this first stage of the Prehistory of Peasant Societies.

Portuguese Beiras and Northern Meseta things are not as clear cut because of the “parasitic” reuse (Jorge, et alii., 1997) of megalithic tombs and continuity of use of natural cavities4 (fig.4).

Antonio Gilman’s idea that “...the intensification of copper age collective burial rites [...] is meant to mediate the incipient social differentiation of the third millennium.” (Gilman, 1987, p. 29) accommodates well the new evidence provided by the proliferation of the recently discovered ditch-limited enclosures as architectural multifunctional expressions within the regional settlement systems. A late and transitional2 example is provided by the small funerary (?) enclosure of Bela Vista 5 (Valera, 2013).

Figure 4.4: Halberd, Palmela point and axe of arsenic copper. Part of the large set of metalic artefacts and pottery vessels from the EBA individual burial from Gruta das Redondas in Central Portugal.

 

Nevertheless, the individual burials from Montelavar (Sintra – Harrison, 1974) and Gruta das Redondas (Carvalhal de Aljubarrota – Natividade, 1901) and other known cases from the Portuguese Beiras, in parallel with the “parasitic” reuse of megalithic monuments – as in the case of the later Beaker burial in the corridor of Orca de Seixas (Senna-Martinez, 1994a) – indicate a tendency to such individual treatment. For Beira Alta, new burial types under small and non-megalithic tumulus are other examples of such a tendency (Cruz, 1998; Cruz, Gomes & Carvalho, 1998a e b).

3. The Bronze Age, between change and continuity 3.1. The First Bronze Age3 Around the beginning of the last quarter of the third millennium BC, changes in several aspects of early Iberian peasant societies can be detected, pertaining mostly to their social organization and symbolic systems, and marking the beginning of the Bronze Age (SennaMartinez, 2009). These are not uniform across the different Iberian regions but, nevertheless, allow us some degree of generalization, namely three discontinuities in the field data materialise these changes: (1) the decline, abandonment and restructuration of settlement systems; (2) the end of megalithic collective burial; and (3), linked to the above mentioned two, the emergence of a new symbolic system is revealed by the fading of feminine iconographic representation, the development of an andriarcal iconography, and the role of metal weapons and jewellery as social markers of prestige and power (Senna-Martinez, 2007, p.120).

In the North of Portugal, namely in Minho, individual treatment of the dead is clear. Examples range from a few exceptional burials – for instance in the case of Quinta da Água Branca cist burial – to simpler situations in what concerns grave goods as in the Vale Ferreiro necropolis (Bettencourt, Lemos e Araújo, 2002; Bettencourt, et al. 2005). In this regional area and besides the generalized individualization of funerary ritual there exists a broad polymorphism of building solutions ranging from “parasitic” reuse of megalithic tombs to the internal variety detected in various necropolises, such as Vale Ferreiro. Once lost the architectural monumentality of the Neolithic megalithic tombs it becomes harder to identify the existence of post-inhumation rituals as their visibility diminishes. Nevertheless, in the Argaric and Southwest cultural areas we already have an abundance of evidence for commensality rituals associated with almost the complete range of burial types (Aranda Jiménez e Esquivel Guerrero, 2006, 2007; Gomes et al. 1986; Gomes, 1994; Alves, et al. 2010). Such a possibility also

The rise of an individual treatment in death as opposed to the megalithic collective rituals is clearly marked in the Peninsula southern half, namely the Southwest, the Argaric areas and Mancha. For the Atlantic Estremadura,

                                                             2

In the sense that it presents both vestiges of the “neolithic” practices together with what we can call an individual “prestige” central pit inhumation of a woman. 3 For the sake of simplicity, we consider in this paper the Bronze Age divided into a First Bronze Age [encompassing the traditional Early (EBA) and Middle Bronze Ages (MBA) – c. 2250-1250 BC] and a Late Bronze Age (LBA – from c. 1250 BC up to regional beginning of the Iron Age – c. 850 BC in the southern half of the Peninsula and c. 450 BC in its Center and NW).

                                                             4

As for instance in the case of the recently revised and republished set of materials from Furnas do Poço Velho (Cascais – cf. Gonçalves, 2009: 135-137).

38 

 

exists, in Northern Portugal, for Monumento 1 of Outeiro de Gregos (Jorge, 1990) with its peripheral structure.

(knives, awls) and also jewellery (mainly in beaten goldplate and rarely in silver – diadems, finger rings, bracelets or dress applications).

In Iberian First Bronze Age another generalized transformation in burial ritual concerns the kind of grave goods that accompany the deceased. While in the Late Neolithic and Chalcolithic the burial space can be seen as an extension of the domestic space – where the range of grave goods replicate the ones used in the living quarters (Senna-Martinez e Ventura, 2008, p.342; SennaMartinez, López Plaza & Hoskin, 1997, p.666) – in the First Bronze Age only a few items were chosen.

The scarcity of the items produced in copper or arsenical copper – we don’t even need to mention the very few items in gold or silver – and their main association as funerary offerings of a select few burials6 since the EBA, all points towards a non-technomical character for this early metallurgy of Iberia (Roberts, 2009, p.472). The EBA is the period when the first situations of metal artefacts ritual deposition as land or water “deposits” – as material sacrifice or “death” of the deposited items – occur (Senna-Martinez, 2009). In almost all the registered occurrences these deposits are constituted of halberds blades (fig.6), rarely with a few other items that can, nevertheless, include a few jewellery items.

The situation is clear cut namely in what concerns pottery where a very limited ranges of vessel types have funerary use and sometimes they are predominantly produced to such use5 (fig.5).

Parallel to their funerary or ritual deposit, these exceptional “packages” of weapons and jewellery items can appear as iconographic depiction, either per si7 or (less often) as paraphernalia of the first “power anthropomorphic representations”  8 that extend into the MBA. From this EBA metal “package” the first long daggers (or tongue-swords) and principally the halberds (fig.6) constitute supra-regional types of artefacts that either as part of “deposits”, a funerary item or an iconographic depiction clearly illustrate this new “discourse of power” (Senna-Martinez, 1994b e 2007). Being a supra-regional symbol, the halberds nevertheless show different regional treatment alongside Western and South Iberia. They seem to function as a way of regional elites being “a la page” while maintaining different regional representational idiosyncrasies. So and without pretending to be exhaustive: (1) Galicia and Minho regions will see “deposits” marking land and water inter-areas points of transit (cf. Leiro, Rianxo – Meijide, 1989) as well iconographic representation in petroglyphs (Costas Goberna, et al. 1997); (2) in Eastern Trás-os-Montes and part of Zamora province we will have only land and water “deposits” (Senna-Martinez, 2006) with stelae close by, as in the cases of Valdefuentes de Sangusín (Salamanca – Santonja Gómez and Santonja Alonso, 1978) and Longroiva (in Beira Transmontana – Almagro, 1966), as two clear “power figures”;

 

Figure 4.5: Tronco-conic vessels. Part of the pottery offerings from the EBA individual burial from Gruta das Redondas. Besides specially produced pottery vessels, the distinctive mark of “elite burials” mainly consists of metal artefacts: (1) weapons for the male ones (halberds, tongue-daggers, Palmela points, tongue-swords, , in the EBA and, in the MBA, axes with large cutting-edge, riveted-hilt swords, riveted daggers, and long tanged points), as well as jewellery (beaten gold diadems, sword-handle decoration, and “archer brassards”); (2) utensils for the female ones

                                                             6

Even in the Argaric Southeast which is the socially most complex area of the Iberian Peninsula (Castro Martínez, et al. 1993-94). 7 This is both the case of the engraved stone schist slabs covering cistburials in the Southwest First Bronze Age (Early and Middle – Barceló, 1991; Gomes & Monteiro, 1977) as well of the Galician petroglyphs (Costas Goberna, et al. 1997). 8 As in the case of some stelae and menhir-statues from the Portuguese Beiras and Trás-os-Montes (Sanches & Jorge, 1987; Jorge & Jorge, 1990) as well as the stelae from the Southwest (Barceló, 1991; Gomes, 1994; Gomes & Monteiro, 1977).

                                                             5

As, for exemple, in the case of the so called Siret’s type 6 (Castro Martínez, et alii., 1993-94: 102), the tronco-conic vessels of Central an Northern Portugal areas (Senna-Martinez, 1993 e 2000: 107) and the “rippenvase” or “zonenvase” from the Southwest (Schubart, 1975: 4649).

39 

 

Figure 4.6: Halberd types distribution from the Iberian Peninsula (cf. Senna-Martinez, 2007, fig.1). (3) in the Southwest (Alentejo and Algarve) cases of a probable land “deposit” (Cano, Sousel – Carreira, 1996) coexist with funerary depositions (Vale de Carvalhos – cf. Arruda et al. 1980; Antas, Tavira – cf. SennaMartinez, et al. 2013), an habitat site (Castillo de Alange – Pavón Soldevila, 1994) and pictographic depictions in stelae and schist slabs to cover inhumations (SennaMartinez, 2007, p.124-5, fig.4); (4) the Southeast (the so called “argaric area”) privileges funerary depositions (Castro Martínez, et al. 1993-94).

Three infant and juvenile burials with metal offerings, one from Peñalosa (Burial 5-CE VIIa – Contreras Cortés, Sánchez Ruiz and Nocete Calvo, Eds. 2000, p.212) and two from Cerro de la Encina (Burials 8 and 10 – Aranda Jiménez and Molina González, 2006, p.52-53 and Table I), the last two including pieces of gold and silver jewelry, together with a clear spatial differentiation of the dead – with elite habitat areas as well as tombs occupying the habitats acropolis – testify to probable hereditarily social status and clear social stratification (Aranda J iménez and Molina González, 2006; Jaramillo Justinico, 2004).9

In Iberia the MBA – c. 1750-1250 BC – will see the disappearance of the halberds as a “symbol of power”, being replaced by flat axes, with large cutting edges in relation to the butt, that in the Iberian Northwest (Galicia, Minho and Trás-os-Montes) are designated as of “Bujões/Barcelos” type (Senna-Martinez, et al. 2013). Found either in “ritual deposits” (from the Northwest to the Estremadura in Central Portugal), or as funerary offerings (mainly in the Argaric area – Castro Martínez, et al. 1993/1994) this occurrence also replaces the halberds by axes as “symbols of power”, that seem to uphold the same symbolic meaning (Senna-Martinez, 2009).

The Argaric situation contrasts strongly with the observed realities of other Iberian regional areas, where infant and juvenile burials are very rare and prestige items conspicuously absent. This situation hints to most of the Iberian regions being organized into lose polities where non-hereditary social status was mostly acquired during each individual lifespan, within probably very small local/regional elites that monopolized production and consumption of scarce prestige items.

                                                             9

This is not the place to a full discussion of the type of political organizations present in the El Argar Culture Group. Nevertheless, we tend to accept the recently expressed position of Aranda and Molina that Argaric polities should be perceived as multiple “... political units of a regional nature controlled by power centres such as Cerro de la Encina, El Cerro de la Virgen or El Argar itself...” (Aranda and Molina, 2006: 58), that we should preferably call “chiefdom”, still close to a “tribal” basis (Fowles, 2002), rather than “state” as pretended by Risch and Lull (1995).

The only regional area in Iberia where we have consistent data hinting to hereditary social status coming into being during the 1stBA is the Argaric Southeast.

40 

 

3.2. The Late Bronze Age. The end of a cycle. The generalization of binary bronze production10 to all Iberia regions is probably one of the turning points at the beginning of the LBA in this peninsula.11 In the funerary realm, nevertheless, metal becomes rarer, accompanying the proliferation of incineration and urn-burial as the preferred way of body disposition. Besides cremation, in the Iberian Southwest, recent research has proved the continuation of pit inhumation as one of the LBA funerary practices (as in Casarão da Mesquita 3 e Monte da Cabida 3 – Soares, et al. 2009) up to now with no funerary goods associated.

 

In the lower Tagus and Atlantic Estremadura region the dominant funerary practice is incineration and urn-burial, as registered in the Tanchoal, Meijão and Cabeço da Bruxa necropolises (Kalb and Höck, 1982 and 1985; Corrêa, 1933-35). Nevertheless, natural cavities continue to be in use as burial ground12 and in its southern extremity exists the only case of reuse of an emptied Chalcolithic tholos as the prestige grave of two male individuals (Spindler, et al. 1973-74; Vilaça and Cunha, 2005; Harrison, 2007).

Figure 4.7: The LBA individual funerary monument of Moinhos de Vento 3 in Central Portugal (SennaMartinez, 1984). Similar is the situation for Northern Portugal and Galicia, with cremation and urn-burial gradually implemented (Bettencourt, 2008). The overall picture for Iberia’s LBA, with cremation becoming the preferred form of body disposal associated to an almost complete lack of burial offerings, suggests the loss of importance of burial as a means of status enhancement and thus its transfer to other forms of representation, namely in the realm of the living. The generalized reconfiguration of settlement systems that occurs in the beginning of the LBA in Iberia opens up new spaces and opportunities of social representation (Jorge, 1995). Of these new opportunities the banquet or “symposium” surely occupied a preeminent place (SennaMartinez, 1996).

With the exception of Roça do Casal do Meio, the absence of prestige grave goods is the rule where the only known cases of metal deposition consist of a very few exemplars of simple bronze bracelets.13 Incineration and urn-burial are also the dominant practices in Central and North Atlantic Iberia with deposition either in small cists or pit. In the first case and in Beira Alta there exist cases of six aggregated cists in a shallow round monument (as in Paranho necropolis – Coelho, 1925) or single cists inserted into a small round stone cairn (for example the Moinhos de Vento 3 and Fonte da Malga monuments – Senna-Martinez, 1984; Kalb and Höck, 1979) or the construction of small megalithic cists inserted in a low cairn (Casinha Derribada – Cruz, Gomes and Carvalho, 1998). In Beira Alta the only case of known metal goods is a bronze bracelet that accompanies the ashes in one of the Paranho cists.

From the communal bearings of the Neolithic under the agricultural cycle as symbolic metaphor, through status and male power figures development during the 1stBA, burial practices of the LBA constitute the end of a cycle that will only be fully transformed again in Iberia when Roman domination will extend urban transformation to all of its regions. But that is another story. References:

                                                            

ALMAGRO, M. (1966) – Las Estelas Decoradas del Suroeste Peninsular. Madrid. Consejo Superior de Investigaciones Científicas (Bibliotheca Praehistorica Hispana, Vol. VIII). ALVES,C.; COSTEIRA, C.; ESTRELA, S.; PORFÍRIO, E.; SERRA, M.; SOARES, A.M.M. and MORENOGARCÍA, M. (2010) – Hipogeus funerários do Bronze Pleno da Torre Velha 3 (Serpa, Portugal). O Sudeste no Sudoeste. Zephyrus. LXVI, p.133-153. ARANDA JIMÉNEZ, G. and MOLINA GONZÁLEZ, F. (2006) – Wealth and Power in the Bronze Age of the South-East of the Iberian Peninsula: The Funerary Record of Cerro de La Encina. Oxford Journal of Archaeology. Oxford. Oxford University Press. 25(1), p.47-59.

10

Binary bronze production generalization can be radiocarbon-dated to the last quarter of the second millennium BC almost everywhere in Iberia (Castro Martínez, Lull and Micó, 1996) namely alongside the Atlantic Facade from the Northwest (Bettencourt, 1999, 2001; Sampaio and Bettencourt, 2011) through the Portuguese Beira Alta (SennaMartinez, et al. 2011), Beira Interior (Vilaça, 1997) to the Southwest (Soares, et al. 2007). 11 In terms of scale of production this is clearly a period where some intensification occurs alongside generalization.11 Nevertheless the generalization of binary bronze production did not significantly change the scale of local ateliers still functioning at a household level and for systems largely of auto-consumption and little circulation (SennaMartinez, 2005). Metal in Iberian LBA was still a luxury prestige item even within more complex social formations and with more typological and formal variety. 12 Sometimes since the Neolithic as in Abrigo Grande das Bocas (Carreira, 1994). 13 Single or in bundles as in Tanchoal (Kalb and Höck, 1995).

41 

 

ARRUDA, A.M.; GONÇALVES, V.S.; GIL, F.B. and FERREIRA, G. (1980) – A Necrópole da Idade do Bronze do Monte de Vale de Carvalho (Sítimos). Lisbon. Clio. 2, p. 59-71. BETTENCOURT, A.M.S. (2008) – Life and Death in the Bronze Age of the NW of Iberian Peninsula. F. FAHLANDER and T. OESTIGAARD, Eds. The Materiality of Death: Bodies, Burials, Beliefs. BAR International Series. 1768. p.99-104. BETTENCOURT, A.M.S. (2001) – Aspectos da metalurgia do bronze no Entre-Douro-e-Minho, no quadro da Proto-História do Noroeste Peninsular. Porto. Arqueologia (GEAP). 26, p. 13-40. BETTENCOURT, A.M.S. (1999) – A paisagem e o homem na bacia do Cávado durante o II e o I milénios AC. Braga. Universidade do Minho. PhD Thesis. 3vols. BOAVENTURA, R. (2009) – As antas e o Megalitismo da região de Lisboa. Lisbon. University of Lisbon. PhD Thesis. CARREIRA, J.R. (1996) – O Conjunto metálico de Cano (Sousel). Aljustrel. Vipasca. 5, p. 59-70. CARREIRA, J.R. (1994) – A Pré-história Recente do Abrigo Grande das Bocas (Rio Maior). Trabalhos de Arqueologia da EAM. Lisboa. Colibri. 2, p.47-144. CASTRO MARTÍNEZ, P. V.; CHAPMAN, R.W.; SURIÑACH, S.G.; LULL, V.; MICÓ PEREZ, R.; RIHUETE HERRADA, C.; RISCH, R. and SANAHUJA YLL, M.E. (1993-94) – Tiempos sociales de los contextos funerarios argáricos. Murcia. AnMurcia. 9-10, p. 77-105. CASTRO MARTÍNEZ, P. V.; LULL, V. and MICÓ, R. (1996.) – Cronología de la Prehistoria Reciente de la Península Ibérica y Baleares (c. 2800-900 BAR Publishing cal ANE). Oxford. (BAR International Series, 652). COELHO, J. (1925) – A Necrópole do Paranho. Viseu. Edição do Autor. CORRÊA, A.M. (1933-35) – Urnenfelder de Alpiarça. Anuario de Prehistoria Madrileña. 4/6, p.133-137. COSTAS GOBERNA, F.J.; HIDALGO CUÑARRO, J.M.; NOVOA ÁLVAREZ, P. and PEÑA SANTOS, A. (1997) – Las representaciones de armas en el grupo galaico de arte rupestre. F.J. COSTAS GOBERNA and J.M. HIDALGO CUÑARRO, Eds. Los Motivos de Fauna y Armas en los Grabados Prehistóricos del Continente Europeo. Vigo. Asociación Arqueológica Viguesa (Serie Arqueología Divulgativa, 3). p. 85-112. CRUZ, D. J.; GOMES, L. F. and CARVALHO, P. S. (1998) – O grupo de tumuli da Casinha Derribada (Concelho de Viseu). Conimbriga. XXXVII, p.5-76. FAIRBAIRN, A.S. (2000) – On the spread of crops across Neolithic Britain, with special reference to the southern England. A.S. FAIRBAIRN, ed. Plants in Neolithic Britain and beyond. Oxford. Oxbow Books, p.107-121. FAIRÉN-JIMÉNEZ, S. (2007) – Rock art and social life: Revisiting the Neolithic transition in Mediterranean Iberia. Journal of Social Archaeology. 7, p.123-143. FOWLES, S.M. (2002) – “From Social Type to Social

Process: Placing ‘Tribe’ in a Historical Framework”. W.A. PARKINSON, Ed. The archaeology of Tribal Societies. Ann Harbour (Michgan). International Monographs in Prehistory. ‘Archaeological Series’. 15, p.13-33. GILMAN, A. (1987) – Unequal development in Copper Age Iberia. E. A. BRUMFIEL and T. K. EARLE, Eds. Specialization Exchange and Complex Societies. Cambridge. Cambridge University Press, p. 22-29. GIMBUTAS, M. (1996) – The Goddesses and Gods of Old Europe. London. Thames and Hudson. GOMES, M.V. (2012) – Early Neolithic Funerary Practices In Castelo Belinho’s Village (Western Algarve, Portugal). J.F.GIBAJA, A.F. CARVALHO and P. CHAMBON, Eds. Funerary Practices in the Iberian Peninsula from the Mesolithic to the Chalcolithic. Oxford. BAR Publishing. BAR International Series. 2417, p.113-123. GOMES, M.V. & MONTEIRO, J.P. (1977) – As estelas decoradas da herdade de Pomar (Ervidel - Beja) Estudo comparado. Setúbal. Junta Distrital de Setúbal. Setúbal Arqueológica. 2/3, p. 281-343. GONÇALVES, V.S. (1989) – Megalitismo e Metalurgia no Alto Algarve Oriental. Uma aproximação integrada. Lisbon. CAH/Uniarch/INIC. ‘Estudos e Memórias’, 2. 2 volumes. GONÇALVES, V.S. (1978) – Para um programa de estudo do Neolítico em Portugal. Zephyrus. Salamanca. 28-29, p.147-162. GONÇALVES, V.S. (2008) – As ocupações préhistóricas das Furnas do Poço Velho (Cascais). Cascais. Câmara Municipal de Cascais. HARRISON, R. J. (2007) – A revision of the later Bronze Age burials from the Roça do Casal do Meio (Calhariz), Portugal. Beyond Stonehenge: Essays on the Bronze Age in Honour of Colin Burgess. Oxford. Oxbow Books. p.65-77. HOSKIN, M. (2001) – Tombs, Temples and their Orientations. Bognor Regis. Ocarina Books. HOSKIN, M., et al. (1998) – Studies in Iberian Archaeoastronomy: (5) Orientations of megalithic Tombs of Northern and Western Iberia. Archaeoastronomy. 23 (JHA, xxix), p. S59-S62. JARAMILLO JUSTINICO, A. (2004) – “Aproxima-ción a la Vida Cotidiana de las Poblaciones Argáricas: El Caso de Peñalosa”. @rqueología y Territorio. 1, p.83-99. JORGE, S. O. (1995) – Introdução. A Idade do Bronze em Portugal, Discursos de Poder. Lisbon. Instituto Português de Museus. p.16-20. JORGE, V. O. e JORGE, S. O. (1990) – Statues-Menhirs et Stèles du Nord du Portugal. Revista da Faculdade de Letras (Porto). II Série. VII, p.299-324. KALB, P. and HÖCK, M. (1982) – Cabeço da Bruxa, Alpiarça, Distrito de Santarém. Portugalia (N.S.), 2/3, p.61-9. KALB, P. and HÖCK, M. (1985) – Cerâmica de Alpiarça. Exposição temporária na Galeria dos Patudos, Câmara Municipal de Alpiarça/Instituto Arqueológico Alemão de Lisboa. KALB, P. & HÖCK, M. (1979) – Escavações na 42 

necrópole de mamoas ‘Fonte da Malga’ - Viseu, Portugal. Beira Alta. Viseu. 38(3), p.593-604. MARTINHO, A.T. (1981) – O Pastoreio e o Queijo da Serra. Lisbon. Parque Natural da Serra da Estrela. 2ª Ed. MEIJIDE, G. (1989) – Un importante conjunto del Bronce Antiguo de Galicia: El depósito de Leiro (Rianxo, A Coruña). Gallaecia. 11, p. 151-164. MONTERO RUIZ, I. (1994) – El Origen de la Metalurgia en el Sureste de la Península Ibérica. Almería. Instituto de Estudios Almerienses. PAVÓN SOLDEVILA, I. (1994) – Aproximación al estudio de la Edad del Bronce en la Cuenca del Guadiana: La Solana del Castillo de Alange (1987). Cáceres. Institución Cultural El Brocense. RISCH, R. e LULL, V. (1995) – “El estado argárico”. Verdolay. 7, p.97-109. ROJO-GUERRA, M.A. and GARRIDO-PENA, R. (2012) – From Pits to Megaliths: Neolithic Burials in the Interior of Iberia. J.F.GIBAJA, A.F. CARVALHO and P. CHAMBON, Eds. Funerary Practices in the Iberian Peninsula from the Mesolithic to the Chalcolithic. Oxford. BAR Publishing. BAR International Series. 2417, p.21-28. ROBERTS, B.W. (2009) – Production Networks and Consumer Choice in the Earliest Metal of Western Europe. Journal of World Prehistory. 22, p. 461-481. SAMPAIO, H.A. and BETTENCOURT, A.M.S. (2011) – Produção e Práticas Metalúrgicas da Idade do Bronze no Noroeste Português: O Caso do Pego, Braga. C.B. MARTINS, A.M. BETTENCOURT, J.I. MARTINS e J. CARVALHO, Eds. Povoamento e Exploração de Recursos Mineiros na Europa Atlântica Ocidental. Braga. CITCEM/ APEQ, p. 391-407. SENNA-MARTINEZ, J. C. (2009) – Armas, lugares e homens: Aspectos das práticas simbólicas na Primeira Idade do Bronze. Estudos Arqueológicos de Oeiras. Oeiras. Câmara Municipal. 17, p. 467-488. SENNA-MARTINEZ, J. C. (2007) – Aspectos e problemas das origens e desenvolvimento da metalurgia do bronze na Fachada Atlântica Peninsular. Estudos Arqueológicos de Oeiras. Oeiras. Câmara Municipal. 15, p. 119-134. SENNA-MARTINEZ, J. C. (2006) – Depósitos metálicos versus economia política das práticas metalúrgicas na Idade do Bronze em Portugal. Comentário a Raquel Vilaça, “Depósitos de Bronze do Território Português. Um debate em aberto”. Lisbon. O Arqueólogo Português. Série IV, 24, p. 109-114. SENNA-MARTINEZ, J. C. (2005) – O outro lado do comércio orientalizante: Aspectos da produção metalúrgica no pólo indígena, o caso das Beiras Portuguesas. Actas del III Simposio Internacional de Arqueología de Mérida: Protohistoria del Mediterráneo Occidental. Mérida. Madrid. CSIC, p.901-910. SENNA-MARTINEZ, J.C. (2000) – O Bronze Pleno. Uma Transformação na Continuidade?. J. C. SENNA-MARTINEZ & I. PEDRO, Eds., Por Terras de Viriato: Arqueologia da Região de Viseu. Viseu.

Governo Civil do Distrito de Viseu e Museu Nacional de Arqueologia, p.105-114. SENNA-MARTINEZ, J.C. (1996) – The symbolism of power in Central Portugal Late Bronze Age Communities. Máthesis. 5, p.163-175. SENNA-MARTINEZ, J.C. (1995) – The Late Prehistory of Central Portugal: a first diachronic view. K.T. LILIOS, Ed. The Origins of Complex Societies in Late Prehistoric Iberia. Ann Harbour (Michigan). International Monographs in Prehistory. ‘Archaeological Series’, 8, p.64-94. SENNA-MARTINEZ, J.C. (1994a) – Notas para o estudo da génese da Idade do Bronze na Beira Alta: o fenómeno campaniforme. Trabalhos de Arqueologia da EAM. Lisbon. Colibri. 2, p.173-200. SENNA-MARTINEZ, J.C. (1994b) – Subsídios para o estudo do Bronze Pleno na Estremadura Atlântica: (1) A alabarda de tipo ‘Atlântico’ do Habitat das Baútas (Amadora). Salamanca. Zephyrus. XLVI, p. 161-182. SENNA-MARTINEZ, J.C. (1993) – Duas contribuições arqueométricas para o estudo do Bronze Antigo/Médio do Centro e Noroeste de Portugal. Trabalhos de Arqueologia da EAM. Lisbon. Colibri. 1, p.77-91. SENNA-MARTINEZ, J.C. (1984) – O Monumento no.3 da necrópole dos Moinhos de Vento, Arganil: a campanha 1(984). Clio/Arqueologia. 1, p.213-216. SENNA-MARTINEZ, J.C.; FIGUEIREDO, E.; ARAÚJO, M.F.; SILVA, R.J.C.; VALÉRIO, P. e VAZ, J. L. I. (2011) – Metallurgy and Society in “Baiões/Santa Luzia” Culture Group: Results of the METABRONZE Project. C.B. MARTINS, A.M. BETTENCOURT, J.I. MARTINS e J. CARVALHO, Eds. Povoamento e Exploração de Recursos Mineiros na Europa Atlântica Ocidental. Braga. CITCEM/ APEQ, p. 409-425. SENNA-MARTINEZ , J. C.; LUÍS, E.; REPREZAS, J.; LOPES, F.; FIGUEIREDO, E.; ARAÚJO, M.F. and SILVA, R.J.C. (2013) – Os Machados Bujões/Barcelos e as Origens da Metalurgia do Bronze na Fachada Atlântica Peninsular. J.M. Arnaud, A. Martins and C. Neves, Eds. Arqueologia em Portugal – 150 Anos. Lisbon. Associação dos Arqueólogos Portugueses, p.599-600. SENNA-MARTINEZ, J.C. e VENTURA, J.M.Q. (2008a) – Neolitização e Megalitismo na Plataforma do Mondego: Algumas Reflexões sobre a Transição Neolítico Antigo/Neolítico Médio. Actas del IV Congreso del Neolítico en la Península Ibérica. Alicante. 2, p. 77-84. SENNA-MARTINEZ, J.C. e VENTURA, J.M.Q. (2008b) – Do mundo das sombras ao mundo dos vivos: Octávio da Veiga Ferreira e o megalitismo da Beira Alta, meio século depois. Homenagem a Octávio da Veiga Ferreira. Estudos Arqueológicos de Oeiras. Oeiras. Câmara Municipal. 16, p.317-350. SENNA-MARTINEZ, J.C. & VENTURA, J.M.Q. (1999) – Evolução das Paisagens Culturais na Plataforma do Mondego na Pré-História Recente (c.5000-550 cal AC). Trabalhos de Arqueologia da EAM. Lisbon. Colibri. 5, p. 9-20. 43 

Funerary Practices in the Iberian Peninsula from the Mesolithic to the Chalcolithic. Oxford. BAR Publishing. BAR International Series. 2417, p.103-112. VALERA, A. C. (2013) – Recintos de Fossos da PréHistória Recente em Portugal: investigação, discursos, salvaguarda e divulgação. Al-Madan. II série. 18, p.93-110 VALERA, A. C. (in press) – Metal, metallurgy, walls and ditches in Portuguese Guadiana basin: an overview. Strategie Insediative e Metallurgia. I Rapporti tra Italia e la Penisola Iberica nel Primo Calcolitico. Roma. Outubro de 2011. VILAÇA, R. (1997) – Metalurgia do Bronze Final da Beira Interior: Revisão dos dados à luz de novos resultados. Viseu. Estudos Pré-Históricos. V, p. 123154. VILAÇA, R.; CUNHA, E. (2005) – A Roça do Casal do Meio (Calhariz, Sesimbra): novos contributos. Almadan. Almada. Centro de Arqueologia de Almada. 2.ª série. 13, p.48-57. WILLIAMS, M. (2003) – Growing metaphors: The agricultural cycle as metaphor in the later prehistoric period of Britain and North-Western Europe. Journal of Social Archaeology. 3, p.223-255. ZAPATA, L.; PENÃ-CHOCARRO, L; PÉREZ-JORDÁ, G. and STIKA, H.-P. (2004) – Early Neolithic Agriculture in the Iberian Peninsula. Journal of World Prehistory. 18(4), p.283-325.

SHERRAT, A. (1981) – Plough and Pastoralism: aspects of the secondary products revolution. I. HODDER, G. ISAAC and N. HAMMOND, Eds. Pattern of the Past. Cambridge. Cambridge University Press, p.261-305. SILVA, A.M.; VALERA, A.C.; LEANDRO, I. and PEREIRA, D. (2012) – Collective cremation burial in Pit 16 from Perdigões Enclosure: a unique funerary context in the Portuguese Chalcolithic burial practices. Poster presented to the 18th Annual Meeting of the European association of Archaeologists, 2012, Helsinki. SILVA, F. (2012) – Landscape and Astronomy in Megalithic Portugal: the Carregal do Sal Nucleus and Star Mountain Range. Papers from the Institute of 22, p.99-114. DOI: Archaeology (PIA). http://dx.doi.org/10.5334/pia.405 SOARES, A.M.M.; SANTOS, F.J.; DEWULF, J.; DEUS, M. and ANTUNES, A. S. (2009) – Práticas rituais no Bronze do Sudoeste: alguns dados. Estudos Arqueológicos de Oeiras. Oeiras. Câmara Municipal. 17, p.433-456. SOARES, A.M.M., VALÉRIO, P., FRADE, J.C., OLIVEIRA, M.J., PATOILO, D., RIBEIRO, I., AREZ, L., SANTOS, F.J.C., ARAÚJO, M.F. (2007) – A Late Bronze Age Stone mould for flat axes from Casarão da Mesquita 3 (São Manços, Évora, Portugal). Proceedings of 2nd International Conference Archaeometallurgy in Europe. Aquileia. Associazione Italiana di Metallurgia. (CD-ROM). SPINDLER, A.; BRANCO, A. C.; ZBYSZEWSKI, G.; FERREIRA, O. V. (1973-74) – Le Monument à Coupole de l’Âge du Bronze final de la Roça do Casal do Meio (Calhariz). Comunicações dos Serviços Geológicos de Portugal. Lisboa. 57, p.91154. VALERA, A. (2012) – Ditches, Pits and Hypogea: New Data and New Problems in South Portugal Late Neolithic and Chalcolithic Practices. J.F.GIBAJA, A.F. CARVALHO and P. CHAMBON, Eds.

44 

The mound at Cimo dos Valeiros (Serra Vermelha, Oleiros, Castelo Branco). A Neolithic burial site in the Central Cordillera, south of Serra da Estrela João Carlos Caninas, Mário Monteiro, André Pereira, Emanuel Carvalho, Francisco Henriques, João Araújo Gomes & Lídia Fernandes, Álvaro Batista

Portugal Central. Por outro lado, alargam a sua distribuição na região de Castelo Branco entre dois extremos altimétricos da bacia do Tejo, ou seja, desde as margens daquele grande rio até aos relevos mais elevados, passando por uma diversidade de orografias intermédias onde se incluem cabeços e lombas do maciço antigo, peneplanícies, mesas detríticas e planícies aluviais.

Abstract Since 2002, archaeological prospection carried out in the context of the Pinhal Interior Wind Farm Project (GENERG Group SA) has led to the discovery and study of mound structures, apparently used for burial purposes, on the final stretch of the Central Cordillera, south of Serra da Estrela.

Palavras-chave: Cordilheira Central (Portugal), Serra Vermelha, mamoa, Neolítico, parque eólico

This paper presents the results of the partial excavation of one of those structures, the mound of Cimo dos Valeiros, located at an altitude of 916 m in the mountains of Serra Vermelha (county of Oleiros). It consists of a mound of around 9 m in diameter constructed in sandy-clay material and capped with clasts of metasedimentary rocks. Excavation revealed an elliptical chamber and archaeological materials that suggest it was in use in the Neolithic.

Introduction Circular artificial mounds in earth and stones were first discovered in the central mountain range (Central Cordillera) of Portugal during the archaeological prospection carried out in the county of Oleiros between 2002 and 2005 in the context of the environmental studies for the GENERG Group Wind Farm Project at Pinhal Interior. These constructions were interpreted as prehistoric burial sites (barrows, tumuli), and the antiquity of some of them (Vale de Mós, Feiteiras and Selada do Cavalo) was confirmed through archaeological excavations. Those studies have since been presented in various scientific conferences and published in specialist journals (Caninas et. al., 2004, 2005, 2008 and 2011).

This data indicates a new boundary in the occurrence of prehistoric burial structures in the uplands of Central Portugal. It also means that their area of distribution in the region of Castelo Branco extends between two high points of the Tagus basin, i.e. from the banks of the river to the highest mountains, passing through various intermediate formations that include hillocks and slopes of the ancient massif, peneplains, detritral tables and alluvial plains.

This paper briefly presents the results of the archaeological intervention carried out at one of these structures, the mound of Feiteiras, now known as Cimo dos Valeiros.

Key Words: Cordillera Central (Portugal), Serra Vermelha, mound, Neolithic, wind farm Resumo

This work was funded by the GENERG Group through the Portuguese archaeological company EMERITA, with logistic and topographical support provided by Oleiros Town Council. The aerial photographs of the tumuli were taken from the top of an aerogenerator by a technician from the company Vestas. IPPAR also supplied one of their technicians (EC) to participate in this intervention. The text was translated by Karen Bennett.

O Projecto Eólico do Pinhal Interior, do Grupo GENERG SA, proporcionou, a partir de 2002, a descoberta e o estudo das primeiras estruturas monticulares, de finalidade aparentemente funerária, em pontos culminantes da Cordilheira Central a Sul da Serra da Estrela. Apresentam-se os resultados da escavação parcial de uma dessas estruturas, a mamoa do Cimo dos Valeiros, situada a 916 m de altitude, na Serra Vermelha (concelho de Oleiros). Consiste num montículo com cerca de 9 m de diâmetro, construído com material areno-argiloso e capeado com clastos de rochas metassedimentares. A escavação revelou uma câmara elíptica e materiais arqueológicos que apontam para uma utilização no Neolítico.

1. Location and background The mound at Cimo dos Valeiros and the two small tumuli nearby (Cimo da Cova dos Bacelos) are located on a long slope that diverges from the dorsal ridge of Serra Vermelha (Figure 5.1). This mountain range, which reaches its maximum altitude at Povoinha, lies in a general NE-SW direction, forming a watershed between the basins of the River Zêzere in the north and the Sertã in the south. It is part of the relief structure that occupies the final stretch of the Portuguese zone of the Iberian

Estes dados indicam uma nova fronteira na ocorrência de arquitecturas funerárias pré-históricas nas terras altas de 45 

 

Central Cordillera, generally known as the Maciço de Alvélos. The slope in question runs between the geodesic point of Povoínha (970 m) and the banks of the River Sertã, underlying the geodesic point of Mosteiro (756 m). The tumuli are set at an altitude of 916 m (Cimo dos Valeiros) and 907 m (the two others), with broad views over the landscape in all directions, except to the north. They are inscribed into the vast formation of metasedimentary rocks of the Beiras Group (Palaeozoic), with occurrences of quartz in philonian seams that fill earlier fractures. The nearest settlement, the village of Cavalo, is on a lower slope, facing westward.

2. Aims The tumuli are located near both a path and the planned line of the electric cable designed to link two of the aerogenerators of the Alvélos Wind Farm. In the wake of the environmental impact assessment, it was decided to keep this cable route away from the tumuli in order to minimize negative impacts of the project. A (partial) archaeological excavation was carried out as a compensatory measure, with the aim of describing those structures and confirming their connection with burial rites in recent prehistory.

Figure 5.1: (A) Location of the Central Inland region of Portugal between the Central Cordillera and the valley of the River Tagus on the relief map of mainland Portugal (adapted from the thematic maps available at www.guiadeportugal.pt). (B) Enlargement of previous map. The square marks the stretch of Serra Vermelha where the tumuli of Cimo dos Valeiros and Cimo da Cova dos Bacelos are located. (C) Location of the tumuli of Cimo dos Valeiros (1) and Cimo da Cova dos Bacelos (2) on an extract of Sheet 277 of the Military Map of Portugal at a scale of 1:25.000 (Army Cartography Department). (D) Location of the tumuli of Cimo dos Valeiros (1), Cimo da Cova dos Bacelos (2) and a rocky outcrop (3) on the topographic plan drawn up prior to the execution of the Alvélos Wind Farm project (source: GENERG).

46 

 

was divided into a grid with cells 1 m2, identified by a letter and a number (A to Q on the xx axis; 1 to 16 on the yy axis). The surface sediment was removed from an area 70 m2, which represented the maximum area achieved by the archaeological excavation. The depth excavation, with dismantling of the structure, covered an area of 13 m2. The sediments removed in both processes were dry-sifted.

In August 2002, when the mound at Cimo dos Valeiros was discovered during the study phase of the wind farm project, there were doubts as to whether it would qualify as a prehistoric burial site, due to the dense vegetation (heather and broom) that covered it. Later, after the vegetation had been cleared away, the archaeological interest of the site was confirmed through surface observation. It consisted of a small semi-circular mound, 8.5 m in diameter, clearly defined above ground, with a dense protective layer of stones (Figure 5.2 - A to D). The depression and rarefaction of the clasts in the central zone seemed to indicate the position of the burial container and the possibility that this had been violated at some point. The centre of the monument was located about 7.5 m west of a path with wheel marks, which followed the line of the ridge. This description was included in the 2005 environmental status report for the wind farm project (RECAPE).

After the SA had been defined, it was found that the centre of the monument was almost perfectly aligned with a linear vertically-stratified outcrop located immediately to the northwest (Figure 5.1 - D3 and Figure 5.2 - C), which we believe would have served as a quarry for the construction of the mound. 3.2. Structure The detailed topographical survey carried out at the start of the intervention (see level curves in Figure 5.4) document a difference in level of around 2 m in the SA with the lowest point located in the southwest corner. The slope of the land, with the accumulation of sediments at the northern side and erosion on the southern side, may explain the deformation of the mound, observed at the surface.

Before 2005, another two barrows had been discovered at Cimo da Cova dos Bacelos, around 120 m south of the mound of Cimo dos Valeiros. They are located a short distance from one another, west of the ridge path mentioned above. The larger mound (Figure 5.3 - G) is around 6 m in diameter and consisted mostly of metasedimentary rock clasts, while the other (Figure 5.3 H), around 4 m in diameter, is mostly of milky quartz and was partially amputated when the path was created. This discovery was made by the archaeologist Carlos Banha of Portuguese Archaeology Institute in the company of two of the authors of this paper (JCC and FH).

Before beginning the intrusive actions of surface sediment removal and excavation, but after the removal of the vegetation from the SA, vertical photographs were taken of each grid cell, taking in the whole of the mound. These were then used to draw Level 1 (surface). The image obtained (partially represented on the western side of Figure 5.4) showed a regular pattern with regard to the distribution of clasts around the central depression, with a greater density on the top. There was also an abnormal accumulation of stones on the northeastern side in cells J6, J7, L6 and L7, which may also have been moved there during a possible violation of site, and a greater scattering of stones on the eastern side. Some larger blocks were integrated into the mound structure or lying loose on top of it, including some which, from their configuration, may have been orthostats, uprooted during the course of the hypothetical violation of the central chamber.

3. Fieldwork The fieldwork, begun in 2006, was directed by JCC and lasted for 33 days. In addition to the authors of this paper, the following archaeologists also participated: Fernando Robles Henriques, José Luis Monteiro, Alexandre Correia, Luis Carvalho, Alexandre Lima, Alexandre Canha and Masters’ student Marta Correia. Although works had been authorized on all three tumuli, the excavation focused exclusively on the mound of Cimo dos Valeiros, due to the limited resources available and also because the work on this monument lasted longer than had initially been anticipated. In the case of the two tombs at Cimo da Cova dos Bacelos, the intervention was restricted to a detailed topographical survey of an area of 15 m x 17 m that included the two structures. The study areas (of Cimo dos Valeiros and Cimo da Cova dos Bacelos) were connected to the geodesic network by Inês Fernandes, topographer at the Oleiros Town Council.

The stone structure visible on the surface consisted mostly of metagraywacke clasts of a range of calibres, from pebbles to boulders (Wentworth scale). There were also clasts of milky quartz, though these were smaller and fewer in number. The proximity of the outcrop, from which elongated boulders still protrude today (Figure 5.2 - E), may have motivated the choice of stone in the construction of this monument, though there might also have been symbolic reasons for aligning the quarry with the mound.14

3.1. Study area The archaeological study area (SA) at Cimo dos Valeiros consisted of a square with sides of 16m, divided by two axes at right angles intersecting at the geometric centre of the monument (depression). One of those axes was oriented in a roughly north-south direction following the gradient of the slope of the ground. The SA (256 m2)

We decided to begin the surface sediment removal in the northeastern quadrant (qNE) as the mound structure was better conserved there. The objective was to describe the

                                                             14

Mound 1 of Selada do Cavalo (Oleiros) is also in a proximity relationship with a linear outcrop. 

47 

 

structure on the top of the mound and reveal other substructures such as internal or peripheral contention rings. This surface sediment removal extended in area into the southeastern quadrant (qSE) and also partially into the northwestern (qNW) and southwestern (qSW) ones as well, in the zone surrounding the central depression.

A second objective consisted of identifying the burial container which, we supposed, would coincide with the depression observed on the surface. The third was to excavate a radial trench on the mound in order to characterize its vertical structuring. This trench was oriented in a south-north direction and was in qNE, in cells L8, M8, N8 and O8. Cell O8 was used for the stratigraphic testing.

Figure 5.2: Views of the monument at the start of work: (A) from the northeast, with the Minas do Cavalo in the background; (B) from the southeast; (C) from the northwest, with the valley of the River Sertã in the background; (D) from the southwest. (E) Alignment of the monument with a 60m-long linear outcrop, with the Serra da Lontreira or Cabeço Rainho in the background. Phases of excavation of the central area, corresponding to a chamber, seen from the south (F); from the east (G); from the north (H).

48 

 

The general surface sediment removal revealed more of the stone structure (SU02)15 in the central and eastern parts of the mound (see partial representation in Level 2 in Figure 5.4). At this stage, it was not possible to know if this structure formed a cairn or if it was simply a superficial protection layer. The stratigraphic unit removed during this process (SU01) consisted of loose dark-brown topsoil (Costa, 2011) that was not very thick and contained ashes and foliage on the top.

successive “levels” of metagraywacke clasts (cobbles and boulders) and the drawing of the respective plans (Figure 5.5 - A). The dimension and arrangement of these elements, which were successively layered to fill the depression, might suggest that they occupied a space that had previously been empty or emptied, and that they had come from a roof consisting of imbricated blocks, possibly supported by wood,16 or had resulted from the fall of the stone roof of the mound into the interior of that depression. Inside the cavity, we found no pieces of a size compatible with orthostats or roof slabs. However, the clasts that we removed could also have resulted from the successive fragmentation of larger blocks, possibly orthostats and pieces of the buttressing around the central container.

The image of Level 2 reinforced what had been obtained in Level 1, indicating the predominance of metagraywacke clasts consisting mostly of cobbles and boulders. The presence of milky quartz, as had been noted on the surface, was minimal and limited to pebbles and cobbles located along the edge of the mound, particularly in the qNE.

This interpretation is supported by the existence, in J9, I8 and I9, of elongated well-entrenched, subvertical blocks tilting inwards towards the interior of the space covered by those cells, and even by J8, suggesting that they might have been primary orthostats, secondary ones or pieces of the buttressing surrounding the central chamber. Moreover, at the limit of the excavation area, and most obviously in the north (in L8 and L9) and west, there were a number of very tilted imbricated slabs, which suggested a buttressing structure for the burial chamber.

On the periphery, particularly in the qNE where the surface sediment removal was more extensive, we found metagraywacke pieces in a horizontal position, disconnected and scattered. Moving closer to the centre of the mound, the lithic fragments became larger and more numerous, and were more tightly imbricated. Many of these pieces were implanted in the ground in both the radial and perimeter positions (in the latter case as if forming a kind of structural brace). In H7, extending into H6 and H8, there was a long thick boulder of metagraywacke, almost 2 m in length, immersed in the tumular mass, in a relatively central position. This was the largest piece found during the excavation. Its function will only be properly understood with more extensive excavations. However, its presence may not necessarily be significant, if we recall the explanation given for the presence of a large stone imbricated in the cairn of Mound 4 of Rapadouro in Vila Nova de Paiva: “the fact that the builders did not make use of it [as an orthostat] was probably due not only to its size (approximately two metres long […]), but also to its irregularity” (Cruz & Canha, 1997: 17).

On the eastern side, particularly in the transition between J7 and J8, there was a hole in the stone covering, which we attribute to a more severe collapse of stones into the central depression. Alternatively, the stones may have been removed during a violation and taken outside and deposited in the adjacent cells to the east, where there was an accumulation of looser stones. To the south, in cells H8 and H9, there was another space that was devoid of any stone structure, possibly as a result of a violation and the removal of the building stones. In support of this hypothesis, there was in cell H7 an elongated block that could have been a small standing stone or a piece of buttressing functioning as a second row of orthostats. Such duplication of the orthostat row is not uncommon in the region (it may be seen in the anta at Silveirinha in Castelo Branco or at the Corgos anta in Idanha-a-Nova, for example) and has been described as “a reinforcement row for the upright slabs of the chamber” (Almeida & Ferreira, 1958).

During this process, a boulder with geometric engravings (Figure 5.3 - E and F) was identified on the southern periphery, in D8. The motif, carved into the end of a metagraywacke boulder, consisted of two open-based triangles in sequence with a slight lateral overlap. The boulder was very fragmented, particularly on the left side above the marks, where there may once have been other engravings. A second stage of work consisted in the excavation of the depression in the central part of the mound, which we believed may have indicated the position of the burial container, given its centrality. Much of the excavation effort was concentrated in this space, with the removal of

                                                             15

  SU02 was defined as the cluster of clasts of various calibres, occupying the whole surface of the mound, which were mostly angular pieces, though also included some rolled pieces, possibly of fluvial origin. They make up the protective stone covering on the surface and the buttresses of the chamber revealed by this excavation. 

                                                             16

  The assessment of the raw material available in the outcrop concluded that it could not have provided a monolithic lid (i.e. a large slab, as occurs with granite architectures).  49 

 

Figure 5.3: Phases of the excavation of the elliptical chamber (A, B). Surface sediment removal, seen from the northwest (C) and north (D). Boulder with carved symbols (E, F). The two tumuli of Cimo da Cova dos Bacelos (G, H), the second amputated by the path on the eastern side. After the clasts of this central area had been removed, the outlines of two depressions were clearly visible (Figure 5.2 - H), the larger one occupying cells I8, I9, J8, J9, L8 and L9 and the other positioned further south in cells H8 and H9. These cavities seem to correspond to a covering over with clay material of two episodes of violation, which may have occurred at the same time or sequentially. The violation seems to have been more severe in the central cavity, given the depth attained, which went beyond the floor of the monument as far as

the geological substratum (SU08). The dimension of the filling suggests that this was a (very) ancient violation. During the excavation, marks were also identified on the edges of small schist clasts, probably made by picks. At the edge of the cavity, particularly on the western and southeastern sides, negative impressions made by stones (orthostats or buttresses) were detected. It was found that the deeply entrenched stone in J9 could not have been a first-line orthostat, as it was set at a relatively high level. 50 

 

The roof of this cavity, now clearly delimited, may have been structured in wood covered by vegetable material and imbricated slabs. The orthostat structure, whose foundations were identified, may have been formed by elongated graywacke blocks extracted from the nearby outcrop, or alternatively, by trunks of wood, though it is not possible to demonstrate this last hypothesis.

However, it could have been a piece of buttressing or a slab from the reinforcement row. On the southeastern side of the cavity, in J8, there was a dense rockfill of large blocks, which meant that it was not permit to detect a possible substructure, such as a corridor, extending in a southeasterly direction. The filling that was removed from this central cavity formed a specific stratigraphic unit (SU04) consisting essentially of clasts of metagraywaque, and some milky quartz. The former were larger and had many crevices filled with a loose dark-coloured sediment with a marked organic component identical to that found in the surface chamber (SU01) (the organic component diminished with depth, with the exception of roots). Underlying this unit and lining the negative impressions of the cavities, we found finer more compact sediment with fewer and smaller clasts (SU3b).

One very interesting aspect of this foundational structure is the northwest-southeasterly orientation of its largest axis, which is not in accordance with the elongation observed in the two violation cavities. On the other hand, this orientation would be compatible with the decentring of this chamber in a northwesterly direction in relation to the stone mass of the mound and its (corrected) geometric centre. This slightly decentred position suggests the extension of the mortuary structure to the southeast, perhaps in the form of a corridor or other type of substructure.

An identical formation with similar characteristics was found at the end of the surface sediment removal process. Later, during the excavation of the trench, this was found to be a structural unit, an embankment (SU03), in a situation of primary deposition, presumably distributed throughout the whole volume of the mound, though this was only partially visible during the removal of surface sediment and excavation of the trench and chamber. It formed the filling of the nucleus of the mound, outside the container and confined between the ground and the protective stone covering. It was around 25 cm thick in the central area, diminishing towards the periphery. SU03b may correspond to the SU03 in secondary position, i.e. the filling of that cavity after it had been abandoned and left to deteriorate (or was violated), with the disappearance of the orthostats.

On the outer edge of SU05 (the sequence of props and cavities that define the contour of the container) we identified a bed of small metagraywacke clasts (SU06), possibly of anthropic origin, underlying SU03. It was marginally identified on the southeastern side of the chamber, although it may have continued in a regular fashion around the whole periphery of this substructure. The third phase of work corresponded to the excavation of the trench in cells L8, M8, N8 and O8. Given the resources available, we aimed to understand the structuring of the mound in depth, which would be easier in this zone as only two cells would have to be excavated (O8 had already been excavated for the stratification test). It would have been very useful if this diagnosis could have been completed with another radial trench (for example, on the boundary between the two easternmost cells), but that was not possible, as already mentioned, due to limited resources.

On the lowest level of the central cavity, there was a compact sedimentary formation (SU07) that was redder in colour, with small metagraywaque plaques, corresponding to the horizon of alteration of the geological substratum. In stratigraphic tests performed on O8, outside the mound, it was found that the surface unit (SU01) passed to the level of alteration of the geological substratum (SU07) through SU03, with a gradual reduction in thickness from the centre to the periphery.

When the stones exposed by the surface sediment removal in M8 and N8 were also removed, it revealed that, at depth (Figure 5.5 - B), there was no continuity with regard to the placement of stones, confirming that this stone structure did not form a cairn but was restricted to a superficial stone covering. Underlying this structure, there was an embankment of light-coloured granulous sediment (SU03), stretching directly down to the geological substratum (SU07, SU08). In cell N8, it was not obvious how the periphery of the mound had been marked, though that limit could have been attributed to the alignment of larger blocks recorded there, thereby conferring a diameter of around 8 m to the mound. For conservation reasons, we decided not to dismantle the structure occupying cell L8, which seemed to include the buttressing of the chamber.

We decided to go ahead with a depth excavation in order to better understand the situation in the central cavity. The results obtained (Figures 5.4 and 5.5 - A) were again very interesting. In fact, the removal of SU03b hid a tenuous though very regular sequence of small embedded clasts and alveoli, configuring SU05, which we interpreted to be the foundation of an orthostat structure since disappeared, which would have been ellipsoidal in contour, with its axis oriented in a northwestsoutheasterly direction.

51 

 

Figure 5.4: General plan of the end of the archaeological intervention, consisting of parts of Levels 1, 2, 8 and 10.

52 

 

Figure 5.5: (A) Final excavation level in the central area of the mound. (B) Cross-section and vertical elevation between the northwestern and northeastern cells, from the centre of AE, including the burial container and trench.

53 

 

1992]) and removed19 by violators. Their absence in the archaeological excavation does not necessarily mean that they were absent from the whole mortuary package.

3.3. Archaeological remains The artefacts of archaeological interest collected during the course of this intervention are represented in Figure 5.6. There were few of them, and they were not very diverse. However, from their type, it is possible to date the use of this structure to the Neolithic period and initiate a discussion about it, considering the presence of significant pieces such as the geometric pieces and blade.

The absence of artefacts within the chamber should be noted. The lack of non-degradable materials inside may have been intentional. In fact, the pieces collected were at superficial levels of the mound, lying there, as (we supposed) they would have been thrown during the course of the violation. The isolation of various pieces and their stratigraphic insertion does not seem to correspond to the ritual deposit on the mound, though they occur in the archaeological record, particularly between the protective covering and the nucleus of the mound (Cruz, 1992: 65).

However, there is not enough data to enable us to fully understand the monument for various reasons. Firstly, these pieces were found on the mound, probably in secondary position and disconnected, given their insertion into SU01, perhaps moved through violation. Though all sediments were sifted, no artefact was found inside the chamber. What is more, the archaeological intervention only partially uncovered the structure of the monument; hence, it would be desirable to continue the investigation, particularly into the easternmost cells, where other substructures and more significant ritual deposits may be hidden.

Alternatively, the lack of artefacts, and even of residues of artefacts, inside that chamber could be explained by their removal at an early stage when the space was less covered over. But we do know of cases of monuments that are empty, such as the anta Amieiro 3 at Idanha-aNova (Cardoso, Caninas & Henriques, 2003), where not only the chamber but also the corridor and atrium were unusually empty.

4. Discussion and conclusions The information that we have obtained from this excavation is still very incomplete from the structural, artefactual and ritual points of view, given the very partial nature of the study. The chamber uncovered could be adjacent to a main chamber that is still hidden. As examples of this scenario, we could cite the western chambers of the tholoi of Lousal 1, Grândola and S. Bartolomé de La Torre, Huelva (Leisner & Leisner, 1959: est. 44 and 49) and Praia das Maças,17 Sintra (Gonçalves, 1983). However, it seems more likely that its decentred position indicates the presence of a corridor running in a southeasterly direction. In accordance with a criterion of presence/absence of certain types of artefacts, our provisional proposal for the chronology of this monument would date its use to the 4th millennium, even taking into account the persistence of geometric pieces and the absence of engraved schist plate and arrow heads. However, this timeframe could only be confirmed with further excavations and absolute datings, without which there is the risk of archaism and isolation due to the mountainous context.

Figure 5.6: Artefacts collected during the archaeological works (indicating the SU and cell position): (1) scraper, silex, SU01, E8; (2) geometric, silex, SU01, H9; (3) geometric, silex, SU01, D7; (4) blade fragment silex, SU01/04, H8; (5) weight or idol, metagraywacke, SU01, L5; (6) mealing stone, metagraywacke, SU01, G2; (7) fragment of the bowl of a hand-made slightly carinated ceramic vessel, SU01, M7; (8) fragment of the bowl of a hand-made ceramic vessel with nipple, SU01, M7.

The absence of polished stone tools (axes, hoes, etc) may be less significant, given that pieces of this type are more easily recognisable and may thus have been coveted (perhaps even for their magical qualities18 [cf. Eliade,

                                                             17

In this site, the remains show that the western chamber corresponds to a first mortuary phase from the Middle-Final Neolithic, while the tholos revealed the use of the Final Neolithic-Early Chalcolithic, according to Gonçalves (1983).  18 As, for example, when they have been inserted into the foundations of houses or into cracks in the walls to protect against electrical discharges during thunderstorms. 

                                                             19

The first regional archaeological chart (Proença Jr, 1910) records only ten polished stone tools in the county of Oleiros. 

54 

 

Figure 5.7: Necropolis at Fonte da Malga (based on Kalb, 1994). However, given the limited nature of this excavation, the hypothesis should be raised that the chamber discovered does not correspond to the burial container, and that this is still hidden. Or, more simply, there may never have existed any tomb at all, and this container is merely a cenotaph.20 In addition, the convergence of structures with depositions alongside others that do not have them is documented in prehistoric mortuary structures (cf. Valera & Filipe, 2012).

as ornamental items and polished stone tools, but with an absence of pottery (Senna-Martinez & Ventura, 2000: 37). This scenario is repeated in Arouca at the Escariz necropolis (Silva, 1988). Amongst the artefacts collected at Cimo dos Valeiros, there is a metagraywacke pebble with roughly hewn side notches (Figure 5.6 – 5), which suggests that it might have been used as a weight, similar to the so-called netsinkers or loom-weights, made from rolled flints which were notched on each side to enable fixation. However, this piece is less symmetrical than those, larger and heavier, and has not one but two pairs of notches on opposite sides.

In fact, an artefact package similar in composition to what we collected can be seen at Anta 4 of Rapadouro in Vila Nova de Paiva, a simple enclosed chamber covered by a 10 m diameter cairn (Cruz & Canha, 1997). This involved six small silex segments, a retouched microblade (hyaline quartz), residual flakestones, also in silex, and some ceramic fragments, forming a group attributed to the second half of the 5th millennium BC. In another mountainous context, there is the case of the Hayas 1 tumulus in Cantabria, attributed to an earlier phase of regional megalithism, with remains restricted to microliths, blades and ceramics (Serna González, 1995).

Its similarity with the anthropomorphic idols of Southeast Iberia and the Mediterranean (Almagro Gorbea, 1973: 55) is very suggestive, if we consider that it is incomplete at the top and possibly also the bottom. In shape, it is also similar to one of the idols collected in the most recent dolmen of the mortuary complex at Dombate, connected to “the world of the south and the Mediterranean” (Bello Diéguez, 1994:289), dating from the 3rd millennium BC.

If we look at the lowlands of southern Portugal, there is the necropolis complex at Atafonas, with its three phases of use (5th and 4th millennia BC), including tombs in the form of pits or with subcircular or oval chambers, covered by tumulus and archaic remains, considered proto-megalithic. The one from the oldest phase (a pit without a tumulus) contained bone remains, “three fragments of handmade ceramics, two large blades and one smaller one (the filling of the pit); two fragments of handmade ceramics, three geometric pieces and possibly one blade (burial level)” (Albergaria, 2007:16).

Other closer parallels may also be suggested. For example, there are the two small pieces identified in sites in the county of Sabugal, alongside manual ceramics and stone-working: a long, rounded pebble, 7 cm in length, with two notches on opposite sides on one of the ends, found in a context attributed to the Bronze Age at the Fornito site and presented as a “figurine in black rock” and “possible pendant” (Marcos, 2012: 54); a fragment of an elongated piece in greenish schist, 4 cm long, considered idoliform, fragmented at the ends and also with two notches on either side at one end, identified at the protohistoric site of Matrena II (Marcos, 2012. 5859).

Artefacts from the oldest phase of the megalithism in the Upper Mondego region, attributable to the Middle Neolithic, are also characterized by “geometric patterns on blades (predominantly crescents and triangles), medium and large blades (mostly not retouched)” as well

The block of metagraywacke engraved with two openbased triangles is another finding of great interest. The representation of broken lines, also known as zigzags, is common in megalithic monuments in Western Europe

                                                             20

The first regional archaeological chart (Proença Jr, 1910) records only ten polished stone tools in the county of Oleiros. 

55 

 

chamber and corridor and on slabs C8 of the chamber, R18 and L12 of the corridor and K2 and K7 of the outer stone ring. These engravings occur at the top of the respective pieces.

and may take the form of engravings (incised or pecked) or paintings. In northwest Iberia, there are cases of obsessive zigzag engravings, both vertical and horizontal, painted on the slabs of the antas of Mota Grande and Portela do Pau 2 (Baptista, 1980), attributable to the second half of the 5th millennium BC, on the anta at Padrão (Cruz & Gonçalves, 1994) in the region of Porto from the first half of the 4th millennium BC, at Pedra Coberta and Castiñeiras 1 and 2 (Peña Santos & Rey Garcia, 1980) in Galicia, or on the Minho coastline on a slab from an anta at Ereira (Silva, 1980) in Afife.

In the Cáceres region of Alcântara, there are numerous examples of such carved shapes, with a strong convergence of zigzags and triangles, both on the slabs of mortuary monuments and on engraved plaques amongst their mortuary remains, with strong connections to anthropomorphic representation. Other cases are found on orthostats in the province of Toledo, on menhirs at Azután and Navalcán, and in Cádiz on the Alberite necropolis (Bueno, et. al., 1999; Bueno Ramírez, 1992; Bueno Ramírez, at al., 1999; Bueno Ramírez & Balbín Behrmann, 2002).

In the Central region, there is the anta of Areita (Gomes, et. al., 1998: 65) in São João da Pesqueira, a monument with a chamber and corridor attributable to the 4th millennium BC, whose head slab has vertical zigzags flanking other geometric figures. One of the zigzags is a short line with three breaks, in the shape of a broken M. And in Spanish Estremadura, four orthostats from the chamber of the tholos at Granja de Toniñuelo (Bueno Ramírez & Balbín Berhmann, 1980) in Badajoz also have broken lines incised or painted, one of them very short. Identical incised zigzags frame the anthropomorphic stele identified there (Bueno Ramírez & Balbín Berhmann, 1980: 114).

One of these cases is the dolmen of Trincones 1, where simple zigzags occur on the orthostats, on a stele from the atrium and on the obverses of two anthropomorphic plaques (Bueno Ramírez, at al., 1999), with a chronology attributed by the authors to the turn of the 4th to the 3rd millennium BC. In the context of the broad representation of triangular figures, and particularly of zigzags in the European megalithic graphic tradition, the presence of figures like those at Cimo dos Valeiros may suggest another link between this monument and this cultural and symbolic universe.

But more interesting because of its greater formal proximity with Cimo dos Valeiros is the case of Cist A at Ínsua (Penedo Romero & Fábregas Valcarce, 1980) in Corunha, dating from the turn of the 3rd to the 2nd millennium BC, which has four orthostats engraved with simple triangular bands, which the authors liken to the graphic representations on the so-called engraved schist plates, aligning them with the megalithic art tradition.

It is open to discussion whether the burial complex of Feiteiras, with its three tumuli, forms a necropolis, indicating some form of continuity with ritual differentiation, or only spatial convergence with cultural discontinuity. Cimo dos Valeiros, larger in size, is around 120 m away from the two monuments of Cimo da Cova dos Bacelos (themselves very close together, with only 8 m between their respective centres). The response must necessarily be connected not only to external morphological aspects, which indicate differentiation or polymorphism, but also to the intrusive study of the two small structures of Cima da Cova dos Bacelos. This external morphological variability occurs in other sites within the county of Oleiros (Caninas, et. al., 2008), where significant variations are found in the size and composition of mound structures, often in a proximity relationship.

In the open air, we could mention the presence of simple zigzags, with three breaks (in M), and open-base triangles, isolated or in pairs, at Pedra Escrita de Ridevides (Santos Jr, 1963) in Alfândega da Fé, a profusely carved rock, which is usually dated to between the Bronze and Iron Ages, the two incised Ms of Alto do Pobral (Caninas, et. al., 2004) in Oleiros. The zigzag with three breaks, in the shape of an M, seems to be individualized in a type relative to the incised carvings of this region (2D form or simple zigzags in Coimbra & Garcês, 2013). We could also cite their occurrence in panel 5 of Cueva del Castillo de Monfrague (Collado Giraldo & Garcia Arranz, 2006), the menhir of Vale de Rodrigo (Gomes, 1994) and the menhir-stele of the enclosure (Gomes, 2000) of Portela de Mogos (stele-menhir 1) and Almendres (stele-menhir 76), attributable to the Final Neolithic.

The polymorphism of burial structures has been recorded and discussed in relation to other necropolises (Jorge, 1981), in some cases with a broad timespan betwen the Neolithic and Bronze Age, as for example at Fonte da Malga (Kalb & Hock, 1982; Kalb, 1994, Figure 5.7) in Viseu, where two Neolithic mounds are found side by side with six cairns from the Final Bronze Age. This necropolis is located in a mountain pass at an altitude of 720 m, crossed by five paths flanked with smaller monuments.

Of similar interest, despite the geographic distance, are similar carvings (in a context marked by numerous triangular figures of various types and zigzags) on orthostats of the monumental mortuary tumulus of Newgrange (O’Kelly, 1982), on the kerb, in the burial

The necropolis of Rapadouro is also formed by four clearly differentiated tumuli (Cruz & Canha, 1997). Three 56 

 

To conclude, the archaeological studies carried out at Cimo dos Valeiros provide results that are of interest for our understanding of the former human presence in the Portuguese zone of the Iberian Central Cordillera, thereby filling a gap in our knowledge resulting from the lack of archaeological investigation in this mountainous area. This possibility is exclusively due to the wind farm projects of the GENERG Group.

of those monuments are closer together, while the fourth is a significant distance away. The first three are attributed to the end of the Chalcolithic/beginning of the Bronze Age, while Monument 4 would have been occupied earlier (Middle Neolithic). A final aspect that deserves attention and to which we will return on another occasion is the alignment between the mound of Cimo dos Valeiros with the linear rocky outcrop. It is of interest to understand if that alignment occurred by chance, though the proximity may have been determined by the fact that it represented a source of raw material (we presume it was used as a kind of quarry), or if the construction was located there for symbolic reasons.

At Cimo dos Valeiros, it was possible to document (more clearly than in Selada do Cavalo 1 or in the tumuli excavated in the counties of Miranda do Corvo [Caninas, et. al., 2012] and Pampilhosa da Serra [Batata & Gaspar, 2009, 2011]) an architecture that dates back to the Neolithic, and which largely exceeds the previous limits of the settlement of that region attributed to the Final Bronze Age (Batata, 2006), taking as reference above all the archaeological materials collected there. The burial structure, which is in keeping with known patterns of “earth” mounds with stone coverings, converges with this chronological attribution.

Other examples of such spatial convergences may be found between megalithic monuments and natural resources. One such case is the 41 tumuli of Escariz (Arouca), many of which are rigorously aligned (in a straight line) with a quartz seam and the frontier between metagraywackes and granites (Silva, 1987: 24).

taxonómicos y cronológicos. Estudos Pré-Históricos. Viseu: CEPBA, nº 2, p. 287-304. BUENO RAMÍREZ, P. (1992) – Les plaques décorées alentéjaines: approche de leur étude et analyse. L’Anthropologie. Paris: S.A.P, t. 96, nº 2-3, p. 573604. BUENO RAMÍREZ, P. & BALBÍN BEHRMANN, R. (1980) – Arte megalítico en sepulcros de falsa cúpula. A propósito del monumento de Granja de Toniñuelo (Badajoz). Brigantium. A Coruña: Boletín do Museu Arqueolóxico e Histórico da Coruña, nº 10, p. 91-121. BUENO RAMÍREZ, P. & BALBÍN BEHRMANN, R. (2002) – L’Art mégalithique péninsulaire et l’Art mégalithique de la façade atlantique: un modèle capillarité appliqué à l’Art post-paléolithique européen. L’Anthropologie. Paris: Elsevier, t. 106, nº 4, p. 603-646. BUENO RAMÍREZ, P.; BALBÍN BEHRMANN, R.; BARROSO BERMEJO, R.; ALDECOA QUINTANA, M. A. & CASADO MATEOS, A. B. (1999) – Arte megalitico en Extremadura: los dólmenes de Alcântara, Cáceres, España. Estudos Pré-Históricos. Viseu: CEPBA, nº 7, p. 85-110. BUENO, P.; BALBÍN, R. DE; BARROSO, R.; ALCOLEIA, J. J., VILLA, R. & MORALEDA, A. (1999) – El dólmen de Navalcán. El poblamiento megalítico en el Guadyerbas. Monografias. Toledo: Instituto Provincial de Investigaciones y Estudios Toledanos, nº 1, 136 p. CANINAS, J. C.; HENRIQUES, F.; BATATA, C. & BATISTA, Á. (2004) – Novos Dados sobre a PréHistória Recente da Beira Interior Sul. Megalitismo e Arte Rupestre no Concelho de Oleiros. Separata da revista Estudos de Castelo Branco. Castelo Branco, Nova Série, 3, 30 p. CANINAS, J. C.; HENRIQUES, F.; BATATA, C.;

References ALBERGARIA, J. (2007) – O sítio neolítico das Atafonas (Torres de Coelheiros, Évora). Revista Portuguesa de Arqueologia. Lisboa: IPA, volume 10, nº 1, p. 5-35. ALMAGRO GORBEA, M. J. (1973) – Los Idolos del Bronce I Hispano. Bibliotheca Praehistorica Hispana. Madrid: Consejo Superior de Investigaciones Científicas, nº 12, 354 p. ALMEIDA, D. F. & FERREIRA, O. DA V. (1958) – Duas sepulturas megalíticas dos arredores de Idanhaa-Velha. Revista de Guimarães. Guimarães: Sociedade Martins Sarmento, nº 68, p. 317-322. BAPTISTA, A. M. (1980) – Arte megalítica no Planalto de Castro Laboreiro (Melgaço, Portugal and Ourense, Galicia). Brigantium. A Coruña: Boletín do Museu Arqueolóxico e Histórico da Coruña, nº 10, p. 191-216. BATATA, C. A. M. (2006) – Idade do Ferro e Romanização ente os rios Zêzere, Tejo e Ocreza. Trabalhos de Arqueologia. Lisboa: IPA, nº 46, 289 p. BATATA, C. & GASPAR, F. (2009) – Mamoas e arte rupestre no concelho de Pampilhosa da Serra (Centro de Portugal). AÇAFA On-line. Vila Velha de Ródão: A.E.A.T, (www.altotejo.org), nº 2, 21 p. BATATA, CARLOS & GASPAR, F. (2011) – Trabalhos arqueológicos desenvolvidos na mamoa VI de Vilares e nos painéis de arte rupestre de Bregada – Unhais-o-Velho, Pampilhosa da Serra. AÇAFA Online. Vila Velha de Ródão: A.E.A.T, (www.altotejo.org), nº 4, 23 p. BELLO DIÉGUEZ, J. M. (1994) – Grabados, pinturas e ídolos en Dombate (Cabana, La Coruña). Grupo de Viseu ou Grupo Norocidental? Aspectos

57 

 

BATISTA, Á.; SABROSA, A.; CANHA, A.; HENRIQUES, F. R.; CHAMBINO, M. & MONTEIRO, M. (2005) – Serra de Alvélos. Sepulturas sob montículo artificial e gravuras rupestres. Catálogo da Exposição 25 Sítios Arqueológicos da Beira Interior. Trancoso: ARA – Associação de Desenvolvimento, Estudo e Defesa do Património da Beira Interior/Câmara Municipal de Trancoso, p. 40-41. CANINAS, J. C.; SABROSA, A.; HENRIQUES, F.; GERMANO, A.; MONTEIRO, J. L.; CARVALHO, E.; BATISTA, Á.; CANHA, A. & CHAMBINO, M. (2011) – Tumulus de Vale de Mós 1 (Serra Vermelha, Oleiros). AÇAFA On-line. Vila Velha de Ródão: A.E.A.T, (www.altotejo.org), nº 4, 51 p. CANINAS, J. C.; SABROSA, A.; HENRIQUES, F.; MONTEIRO, J. L.; CARVALHO, E.; BATISTA, Á.; CHAMBINO, M.; HENRIQUES, F. R.; MONTEIRO, M.; CANHA, A.; CARVALHO, L. & GERMANO, A. (2008) – Tombs and rock carvings in the Serra Vermelha and Serra de Alvélos (Oleiros – Castelo Branco). In PRIMITIVA BUENORAMÍREZ, ROSA BARROSO-BERMEJO & RODRIGO DE BALBÍN-BERHMANN (coord.) – Graphical Markers and Megalith Builders in the International Tagus, Iberian Peninsula. Oxford: BAR Publishing, BAR International Series 1765, p. 89-102. CANINAS, J. C.; MONTEIRO, M.; PEREIRA, A.; CUNHA, P. P.; CARVALHO, E.; LIMA, A.; HENRIQUES, F.; FERNANDES, L. & VIEIRA, M. (2012) – Intervenção geo-arqueológica na mamoa do Penedinho Branco (Vila Nova, Miranda do Corvo, Serra da Lousã). Olhares sobre a Geologia e a Arqueologia de Vila Nova, de Miranda do Corvo e da Serra da Lousã. Vila Nova: Junta de Freguesia de Vila Nova, encontro em Outubro de 2010, p. 40-63. CARDOSO, J. L.; CANINAS, J. C. & HENRIQUES, F. (2003) – Investigações recentes do megalitismo funerário na região do Tejo Internacional (Idanha-aNova). O Arqueólogo Português. Lisboa: Museu Nacional de Arqueologia, nova série, 21, p. 151-207. COLLADO GIRALDO, H. & GARCÍA ARRANZ, J. J. (2006) – La Cueva del Castillo de Moanfrague. Guías Arqueológicas de Extremadura. Mérida: Junta de Extremadura, 21 p. COIMBRA, F. & GARCÊS, S. (2013) – Arte rupestre incisa entre o Tejo e o Zêzere: contributo para o seu inventário, tipologia e datação. Arkeos, perspectivas em diálogo. Tomar: CEIPHAR, nº 34, p. 243-254. COSTA, J. B. DA (2011) – Caracterização e constituição do solo. Manuais Universitários. Lisboa: Fundação Calouste Gulbenkian, 527p. CRUZ, D. (1992) – A mamoa 1 de Chã de Carvalhal (Serra da Aboboreira). Coimbra: Instituto de Arqueologia da Faculdade de Letras de Coimbra, Anexos de Conimbriga, nº 1, 168 p. CRUZ, D. & CANHA, A. (1997) – Escavação arqueológica da mamoa 4 do Rapadouro (Pendilhe, Vila Nova de Paiva, Viseu). Conímbriga. Coimbra: Instituto de Arqueologia da Faculdade de Letras de Coimbra, nº 36 p. 5-26. 58 

CRUZ, D. & GONÇALVES, A. H. B. (1994) – Novas pinturas no dólmen do Padrão (Baltar, Paredes, Porto). Estudos Pré-Históricos. Viseu: CEPBA, nº 2, p. 383-393. ELIADE, M. (1992) – Tratado de História das Religiões. Lisboa: Edições Asa, 572p. GOMES, L. F. C.; CARVALHO, P. S. DE; PERPÉTUO, J. M. A. & MARRAFA, C. (1998) – O Dólmen de Areita (S. João da Pesqueira, Viseu). Estudos PréHistóricos. Viseu: CEPBA, nº 6, p. 33-93. GOMES, M. V. (1994) – Menires e cromeleques no complexo cultural megalítico português – trabalhos recentes e estado da questão. Estudos Pré-Históricos. Viseu: CEPBA, nº 2, p. 317-342. GOMES, M. V. (2000) – Estátuas-menires antropomórficas do Alto Alentejo. Descobertas recentes e problemática. Brigantium, 10. A Coruña: 255-279. GONÇALVES, J. L. M. (1983) – Monumento préhistórico da Praia das Maças (Sintra). Notícia preliminar. Sintra: Sintria, 1-2, p. 29-58. JORGE, V. O. (1981) – Importância do núcleo megalítico de Outeiro de Gregos (Serra da Aboboreira, Baião). Arqueologia. Porto: GEAP, nº 3, p. 29-35. KALB, P. & HOCH, M. (1982) – Escavações na necrópole de mamoas “Fonte da Malga” – Viseu, Portugal. Separata da revista Beira Alta. Viseu: Assembleia Distrital, nº 38 (3), 12 p. KALB, P. (1994) – Reflexões sobre a utilização de necrópoles megalíticas na Idade do Bronze. Estudos Pré-Históricos. Viseu: CEPBA, nº 2, p. 415-426. LEISNER, G. & LEISNER, V. (1959) – Die Megalithgraber Der Iberischen Halbinsel, Der Westen. Berlin: Walter de Gruyter, 349 p. MARCOS, P. NAVE (2012) – Levantamento dos sítios arqueológicos de Aldeia da Ponte. Sabucale. Sabugal: Museu Municipal, nº 4, p. 45-70. O’KELLY, M. J. (1994) – Newgrange, archaeology, art and legend. London: Thames and Hudson, 250 p. PEÑA SANTOS, A. & REY GARCIA, J. M. (1980) – Arte parietal megalítico y Grupo Galaico de arte rupestre: una revisión crítica de sus encuentros y desencuentros en la bibliografía arqueológica. Brigantium. A Coruña: Boletín do Museu Arqueolóxico e Histórico da Coruña, nº 10, p. 301331. PEÑALVER, X. (2005) – Los crómelech pirenaicos. Bolskan. Huesca: Instituto de Estudios Altoaragoneses, nº 22, 349 p. PENEDO ROMERO, R & FÁBREGAS VALCARCE, R. (1980) – Cistas decoradas de Galicia y su contexto regional. Brigantium. A Coruña: Boletín do Museu Arqueolóxico e Histórico da Coruña, nº 10, p. 333342. PROENÇA JR, F. TAVARES DE (1910) – Archeologia do districto de Castello Branco: 1ª contribuição para o seu estudo. Leiria, 25 p. SANTOS JR, J. R. DOS (1963) – As gravuras litotrípticas de Ridevides (Vilariça). Trabalhos de Antropologia e Etnologia. Porto: SPAE, volume 19, nº 2, p. 111-144. SENNA-MARTINEZ, J. C. & VENTURA, J. M. Q.

(2000) – Os primeiros construtores de megálitos. Catálogo da Exposição Por Terras de Viriato. Viseu: Arqueologia da Região de Viseu, p. 35-38. SERNA GONZÁLEZ, M. R. (1995) – La estacion de Alto de Guriezo – Hayas y el megalitismo en la zona oriental de Cantabria. Cuadernos de Sección. Prehistoria – Arqueologia. Donostia:  Sociedad de Estudios Vascos, nº 6, p. 121-134. SILVA, E. J. L. (1980) – Arte megalítica da costa norte de Portugal. Brigantium. A Coruña: Boletín do Museu Arqueolóxico e Histórico da Coruña, nº 10, p. 179-189. SILVA, F. A. P. (1987) – Características do megalitismo na freguesia de Escariz (Concelho de Arouca). Actas das I Jornadas de História e Arqueologia do Concelho de Arouca. Arouca: Câmara Municipal, p. 21-38. SILVA, F. A. P. (1988) – A mamoa 4 da Aliviada, Escariz - Arouca. Trabalhos de Antropologia e Etnologia. Porto: SPAE, volume 28, nº 1-2, p. 137149. VALERA, A. C. & VICTOR FILIPE, V. (2012) – A necrópole de hipogeus do Neolítico Final do Outeiro Alto 2 (Serpa). Apontamentos de Arqueologia e Património. Lisboa: ERA, arqueologia, (www.nia-era.org), nº 8, 2941.

59 

 

60 

 

Caves, Megalithism and Tumuli – Three diachronic realities in funerary archaeography from Alto Ribatejo – Ana Cruz, Ana Graça, Luiz Oosterbeek

Abstract

Zêzere River

North Ribatejo is a geographic region in which three different geomorphological units meet. This reality provides to the Holocene communities different approaches to death. Within these units are karstic caves in the western ridge, where the Nabão River flows, menhirs and dolmens in the old massif where the Zêzere River flows. We aim to present the different modalities of dealing and of approaching death in a diachronic period since Middle Neolithic to the late Bronze age.

Zêzere River is born in Serra da Estrela, has a 214 kilometre extension and is a Tagus tributary in Constância village. It has a medium flow running through its rocky margins with a fairly steep slope and is placed in the Iberian Central Zone of the Old or Hesperic Massif within our subject geographically area. The dominant lithography is the argillaceous schist’s, gneisses, greywacke and quartzite from Pre-Cambrian formation; interspersed by eruptive rocks and punctually covered by Mio-Pliocene formations (Félix, 1993: 242; Ferreira, n.d.: 37).

Keywords: Death, Caves, Megalithism, Tumuli Resumo O Alto Ribatejo é uma zona geográfica onde convergem três tipos diferentes de unidades geomorfológicas. Essa realidade favorece diferentes abordagens das comunidades holocénicas ao fenómeno da morte. Dispomos de cavidades cársicas na Bordadura Ocidental onde se localiza o rio Nabão e de monumentos dolménicos e meníricos no Maciço Antigo banhado pelo rio Zêzere. Este artigo pretende apresentar estas diversas modalidades de tratamento e de abordagem à morte num período cronológico diacrónico que medeia o Neolítico Antigo e a Idade do Bronze Final. Palavras-Chave: Morte, Grutas, Megalitismo, Tumuli The Geomorphological Framework Nabão River The Nabão River is a Zêzere River tributary, born in Ansião in Sicó and Lousã mountain chains and disembogues in Zêzere River margin. It flows from North to South. During its extension of 66 kilometres and along its course crosses the Western Rim/Western Boundary, area of karstic cavities with Holocene human occupation. This area is part of the carbon sedimentary rocks formed in Jurassic and Cretaceous in Mesozoic era, where relief is divided in elevations and middle height hills ripped by river valleys, sometimes steeped. The Canteirões area, more precisely, is a meander type formation, where the Nabão river valley dovetails, formed in the Jurassic period (Félix, 1993: 241). Its genesis assigns characteristics, such as poor agricultural soil, F class and a lack of humidity typically from the limestone soils. The rocks in the area are mainly limestones, marls, argillaceous sandstones and kaolinitic sandstones (Ferreira, n.d.: 49).

Figure 6.1: Localization of rivers Nabão and Zêzere within the Iberia Peninsula hydrography. Ritualizing gestures supported by archaeography Nabão River valley Therefore, is in a very deep meander of the Nabão River, known by the local populations as Canteirões that we will find the karstic cavities now presented: Gruta do Cadaval, Gruta dos Ossos, Gruta de Nª. Srª das Lapas e Gruta do Morgado Superior. Gruta do Cadaval is a horizontal cavity open in the pinkish, compact, with crystalline aspect limestone. It has two contiguous rooms, both used as burial areas during Holocene.

61 

 

Figure 6.2: Localization of the Canteirões do Nabão karstic cavities. 1. Gruta do Caldeirão, 2. Gruta do Cadaval, 3. Gruta dos Ossos, 4. Gruta de Nª. Srª. das Lapas, 5. Gruta do Morgado Superior. West, the longitudinal band named with J letter and, to South, the transverse band designated by 29, therefore we may deduce a complete separation between the two room burial areas, despite their proximity and material resemblance.

The room 1 has a Norwest-southeast orientation and, in it, collective burials were found without any anatomic connection. The bones collected in the C layer were unstructured emerging dispersal by an area of about 4 square meters, without any trace of a burial pit formation. This layer has a date between V and IV millennium B.C.E. In it were gathered some dozens of vases, including fluted decorated ceramics, ceramics with incision marks decoration, with linear motifs and zigzag, indented rims in high necked recipients and carinated vases. The lithic industry included polished axes with oval and trapezoidal section, a weaving piece, blades and bladelets without retouches and crescent form microliths.

The room 2 is oriented from Southwest to Northeast and was possible to recover a rocky structure, in limestone, with sub-squared plan, without cover, where the human remains were gathered (Cruz, Oosterbeek, 1985; Oosterbeek, 1987a; Oosterbeek, Cruz, 1990; Oosterbeek, 1997; Cruz, 2011). Gruta dos Ossos is a horizontal cavity open in the Jurassic limestone, in the Nabão river right margin that sheltered a secondary inhumation ossuary having one particularity from the bones arrangement point of view. It has a date placing its utilization in the IV millennium B.C.E. Resuming, were detected three moments in the ritual operatory chain: I – inhumation near the cavity back wall at Northeast; II – the bones were uplift after being fleshless; III – straighten the rocky soil over the level 4, near the entrance. This ritual gestures took place in a not churned area, where was possible to find a depositional stratigraphy of bone types, being, in the base

The D layer has provided an individual burial and, also in this case, any kind of built structure was found, instead was observed the use of the stalagmite formation, separating the two rooms, to place the body ending up by being covered with a big fallen block. This layer has a date placing the grave in V millennium B.C.E. It was associated to more than a dozen vases, including some with decoration within Early Neolithic patterns, together with polished stone tools of circular or oval section, blades and bladelets without retouches and trapezoidal microliths. The burial area does not exceed in plan, to

62 

 

Figure 6.3: Plan from the surface of the excavation area from Gruta do Cadaval. the pelvis area (sacrum and iliac bones) covered with sediment, over whose were placed the femurs, tibias, fibulas, patella, calcaneus, astragal and metatarsal also covered with sediment, a third deposition with humerus, radius and ulna, metacarpals, ribs and vertebrae covered with earth and, in top, crowning the ossuary and forming a semi-lunar shape, the seven skulls in the Southwest sector, not covered with sediment. These depositions are crossed by an amount of fallen rocks. It was possible to draw a limit in the ossuary at south, west and northwest, being the east limit defined by the wall and the northeast the zone over the wall. The same limits were marked, in irregular shape, by big blocks.

Gruta de Nª. Srª. das Lapas is in Nabão river right margin and is a small cavity roughly orientated at Northeast/Southwest, with the entrance towards Southeast. We have observed 3 occupation moments: 1 – the most recent, between III and II millennium B.C.E., defined with typological criteria (bell-beaker ceramic and a copper flat axe); 2 – the V, IV millennium B.C.E. occupation exists in a pit burial limited by limestone blocks, lately dated by radiocarbon, having at least two vases, twelve bone discoid (rounded) necklace beads, eleven necklace beads made of an undetermined shell and one perforated Pecten maximus; 3 – the VI, V millennium B.C.E. occupation was identified through the excavation of a stony circular structure from an individual burial at Northwest. The lithic industry consists in lithic tools such as scrappers, burins, tips, blades, bladelets and occasionally nucleus of macrolithic industry (Oosterbeek, 1993b; Oosterbeek, 1997; Cruz, 1997; Cruz, 2011).

Any cover was put over the bone elements, once the superior part of the materials and the top set of the layer, testimonies a long exposition to the stalactites precipitations. We must highlight the possible existence of an individual inhumation given the partial presence of an inferior member in anatomical connection – patella, tibia, fibula and part of the metatarsals.

Gruta do Morgado Superior is in Sicó limestone formations and oolitic limestone of Santo António e Candeeiros. It borders the Fetal stream in its valley marginal limestone walls. It was possible, until now, to define two big human bones concentrations in what we call the burial pit 1 and the burial pit 2. Despite some fallen blocks within the sediment there isn’t any stone built structure defined. We think that a pit was open in the sediment each time an individual was to be buried, does emerging the bones without anatomical connection

The non osteological materials are rare: friable clay body ceramics, with a unique rim fragment and fluted decorated ceramic; lithic industry represented by flint blades and bladelets, microliths, polished stone axes with rectangular section, an oblong triangle and bone industry (Oosterbeek, 1987b; Oosterbeek, 1993a; Oosterbeek, 1997; Cruz, 1997; Cruz, 2011).

63 

 

Figure 6.4: Plan from the base of the excavation of Gruta dos Ossos.

Figure 6.5: Plan of Gruta de Nª. Srª. das Lapas. 64 

 

.

Figure 6.6: Plan of Gruta do Morgado. and any type of bone stratigraphical organization. We discard the possibility of being a second inhumation ossuary once we have bones that seldom occur in similar situations. It gives the idea that the bodies were merely placed in the “pit” (as well as votive materials) without any body specific preparation or process of moving the body after decomposition. The evidences come from the presence of bones like sesamoid and pisiform, minimal size bones that wouldn’t appear in case of bones translation. Thus, without consolidate stone structures and a plan, strikes us as a burial “pit” (this structure doesn’t have a defined area and plan) similar to the C layer burial “pit” of Gruta do CadavalWe have exhumed diverse artefacts since flint arrowheads, arsenical copper arrowhead, to ceramic, adornments, bone industry and mobile art (a lagomorph shaped statuette), typically Early Calcolithic and Early Bronze Age materials. Widening our perspectives beyond these specific area we can compare it to other cavities in Portuguese Estremadura region, like, for example, Lugar do Canto (Ferreira e Leitão, 1981), Lapa da Galinha and Lapa dos Carrascos (Gonçalves, 1978), and also the lower Zêzere valley studied Megalithism. The Canteirões do Nabão area strikes, thus, as involved in the convergence process

between the caves traditions and the dolmens tradition, already recorded in other limestone areas, as Alvaiázere or Torres Novas (Oosterbeek, 1994; Cruz, 1997; Cruz, 2011; Cruz, Graça, Oosterbeek, Almeida, Delfino, 2013). Indeed, we have been able to demonstrate that the Neolithization in Alto Ribatejo follows two originally autonomous processes yet related: a cave Neolithic, placeable in the Mediterranean tradition Early Neolithic, with impressed ceramic, and other Neolithic, which we are going to refer below, and were we lay the Megalithism origins. The works developed since 2010 allowed the suggestion that the karstic region tradition would have kept “Neolithic” (changing thus to the Bronze Age) and the megalithic tradition had evolved to full Calcolithic. The Rego da Murta megalithic set, in Alvaiázere (also in Nabão river basin), installed in the karstic landscape, would mark in a later moment the convergence between the two traditions, assuming the megalithic archetype. However Gruta do Morgado excavations demonstrate that the same phenomenon occurred in the cave burial contexts, what not implying a change in the global interpretation model turns it more complex and diversified.

65 

 

Figure 6.7: Megalithic monuments plant topographically. 1. Pedras Negras Dolmen; 2. Dolmen 1 de Vale da Laje; 3. Pedra da Encavalada; 4. Dolmens from Jogada; 5. Dolmens from Vale Chãos; 6. Alqueidão Orthostats.

Dolmen 1 de Val da Laje is placed in a central hill, visually dominant over the Vale do Casalinho small village, as well as the Zêzere river valley. This monument was well preserved, with the tumulus almost intact. The chamber was churned; the top flat capstone (table) was fallen over one of its prop stones (orthostat). The monument is roughly turned to East, surrounded by a barrow with a gap of 150 to 160 centimetres, from centre to periphery.

diameter. The tumulus is built with sediment and greywacke slabs, and still quartz and quartzite pebbles. It was possible to identify several distinctive monument structures constructive phases: 1st phase – D layer, beginning of III millennium B.C.E.; 2nd phase – C layer, beginning of IV millennium B.C.E.; 3rd phase – C layer top, first third of III millennium B.C.E.; 4th phase – C layer top, first third of III millennium B.C.E.; 5th phase – B layer, last third of III millennium B.C.E.; 6th phase – B layer top, final of III millennium B.C.E.; 7th phase – A layer, first third of II millennium B.C.E.

The mound is cut by paths, at north and south, being at origin probably sub-circular, with about 16 meters

The tumulus excavation revealed two lithic rings of quartz pebbles. So, the monument seems to be formed,

Zêzere River valley

66 

 

Figure 6.8: Plan of the top of dolmen 1 de Vale da Laje. Phase 1 – structure I – monument build directly over the bedrock with a single chamber in a triangle-oval plan and entrance towards Southeast; it is formed by a rounded orthostat at West side being the big bedrock its eastern complement. The capstone (table), fragmented, named the monument as “the mounted stone”. At south, there is a small closing slab. If there was, in the north side, any slab closing the chamber, it was gone. It was found very churned, being the gathered materials, mostly, potter’s wheel made ceramics; Phase 2 – structure III – building of the circular monument addorsed to the big bedrock block, at structure I east side. This structure is composed of big orthostats fallen as “dominos”; in an about ten meter diameter area. These orthostats, as well as the structure I lateral stay, were found placed transversally occupying the entire east front. It had circular plan, but due to post-depositional phenomenon during centuries it was completely churned. it is used the bedrock base to support the stays in transversal position, what may be attributed for a logical, practical or empirical reason to the fact the extreme difficulty of raising an artificial platform in the terrain steep slope; Phase 3 – structure II – stone limited pit burials structures, whose plans have sub circular/oval shapes with a headboard never over two meter square area, in all the structure I mound perimeter, interrupted by structure III. It corresponds to what commonly in a dolmen is classified as barrow having the same functions in terms of sustaining the monolith slabs (Cruz, Oosterbeek, 1998; Cruz, 1997; Cruz, 2011).

from centre to periphery, by the dolmen, followed by a first lithic ring about two metres from the dolmen, also followed by a second lithic ring in periphery and a small earth mound and greywacke slabs (this arranged in stairs) covering all structure (tumulus), meanwhile fallen and partially eroded (Oosterbeek, 1997; Oosterbeek, Cruz, Félix, 1992; Drewett, Oosterbeek, Cruz, Félix, 1992; Cruz, 1997; Cruz, 2011). Pedra da Encavalada is an “atypical” megalithic monument, in the Aldeia do Mato stream left margin, a Zêzere river tributary and implanted almost facing Val da Laje dolmen. Reflecting over the paleoenvironmental data analysis obtained, for the moment, to this period (IV – III millennium B.C.E.), is observed the Arbustus unedo dominance associated with Ericaceae gender and few Quercus, Pinus, Alnus and Olea europaea (ALLUÉ, 2000), and combined with cereals presence. It is built at a hillside with the Old Massif (gneiss) lithological landscape as background, with very steep slopes, ending suddenly in Zêzere River. This monument outstanding characteristic is without doubt, the fact of; its builders have recycled also to funerary purposes the big bedrock block visible in the landscape, eventually a benchmark of the communities. In architectonic terms was possible to distinguish three of the monument constructive-occupation phases, which might, however, be coeval: Phase 0 – gneiss natural bedrock – stone quarry and subsequent monolith extraction; 67 

 

Figure 6.9: Pedra da Encavalada stratigraphic profiles, where we can see the tree monument composing structures.

Figure 6.10: Souto necropolis location. 1. Souto Tumulus 1; 2. Souto Tumulus 2; 3. Souto Tumulus 3; 4. Souto Tumulus 4; 5. Souto Tumulus 5; 6. Porto Escuro Tumulus. Souto necropolis is formed by five tumuli located nearby each other. It is word mentioning that Souto 1 tumulus; the only preserved in good conditions and that provided a cinerary urn, is at six meters from Souto 2 tumulus. This had a substantial difference from the first. Instead of have an urn placed in a pit at centre and covered by pebbles; it had a central square plan cist composed by granite fragments and, we believe, that inside of it must have been placed an also incineration urn. This assumption is

produced by the fact of having found some manually made ceramic fragments inside the cist. The 3, 4 and 6 tumuli are located about 100 meters distant from those and were built side by side with each other, becoming difficult to evaluate its effective diameter. Although excavated until the geological base only the tumulus 3 had some small manually made ceramic fragments. So, we may only draw a ritual operatory chain style for Souto tumulus 1 construction. 68 

 

Figure 6.11: Plan of the Souto tumulus 1 surface. The Souto 1 tumulus is a monument with a circular plan with around six meters diameter. It is built by a stone shell (mound) of different size pebbles of quartz and quartzite. From the funerary structure excavation records we think it is possible reconstitute a proposal of the tumulus ritual construction operatory chain. The first step would have been the gather and storage of raw material to the mound. Next, the soil must have been clean and straighten in an area of about six meter diameter where a central pit had to be open to put the urn; a third step would be to carry the funerary urn containing the cinerary remains with the respective votive offerings. The fourth step must have been to place the urn within the open pit, with a schist slab placed slightly at northeast. The urn rim was surrounded by pebbles and finally the mound was built hiding the urn with the mound of quartz and quartzite pebbles of about ten centimetres high in which where randomly placed small ceramic vases. The mound mustn’t have been covered with sediment becoming a marker to the coeval communities (Cruz, 2011).

and dolmens as burial places have had tendency towards a pastoral economy, using the agricultural system as subsidiary. Although we ought to not undervalue the presence in Early Neolithic of individual burials as is Gruta de Nª. Srª. das Lapas case. The fauna and botanical data from archaeological layers confirms clearly the presence of domesticated fauna in caves Early Neolithic contexts (Bos, Sus, Capra, Ovis) followed apparently by untamed fauna dominance (particularly Cervidae), resuming up and generalizing the domestic fauna to common use (mainly capra) in Late Neolithic or Calcolithic (Allué, 2000; Almeida, 2010). There aren’t clear markers of cereals and other plants domestication in early Neolithic phases. This entire context suggests communities in sedentarization process more based in gather, in hunt and, so, with a high degree of mobility. On one hand, through Portuguese Bronze Age and according to the diverse funerary solutions, as the Souto tumuli cases (in this later phase chronological period), we will have equal matches in diversely societies according with their funerary operatory chains.

Data discussion A problem that can be raised at social level, relates with the question: does each type of funerary modality corresponds to a different society?

If, on other side, the option is to accept a common model at identity and social level, the perspective about the different ways of dealing with death seems to be pragmatic regionalisms adapted to the local geomorphology.

If we accept this question as a statement than we shall see through the V, IV and III millennium B.C.E. a change in the burial rituals either in Nabão river valley, either in the Zêzere river valley megalithic monuments, what makes us think in different types of burials, so different types of lives, however we have to consider it in a wider agropastoral plan. Perhaps these communities that used caves

Does the use of the same burial space, or the same funerary modality, implies the existence of a dominant ideology reflected in the social and economic structure? 69 

 

Or otherwise, what seems to be continuity or a rupture reflex, hides’ big changes in the communities?

collectivization pattern and, again, it becomes individual, as seen in Gruta de Nª. Srª. das Lapas.

In recent prehistory research there are no written fonts what increase the difficulty in interpreting the funerary ideal, so, the characteristic feeling of a community towards death is only shown by its materialities; other manifestations are kept from us like music instruments, costume, dance, and culinary, ritual languages. However, that material culture and architectures are provided with a sense which we may call cultural emotional aspect (Cerrillo-Cuenca, 1990: 190). It will be difficult to isolate, materially speaking, the concept itself of religious phenomenon once it concerns to abstraction, conceptualization and organization (idem, ibid.em). We will try to discuss these assumptions through materialities (records), on one hand, the funerary places and, on the other, the votive objects.

The megalithic funerary rituals as observed in Dolmen 1 de Val da Laje and in the monument recuperation and reuse not only in Neolithic and Calcolithic but also in Early Bronze Age where, in this case, we have an individual burial pit in the short corridor and other in the chamber. The Souto funerary rituals imply a ritual operatory chains process. Part of them, however, and for that reason we attain ourselves to those we can identify, are not visible in the archaeological record. In any of the three cases, the settlements are not near the burial places and, the argument of few systematic prospection is not valid, once this are very well known areas, what does mean that Hodder argument is not applied to our particular regional case.

The space, selected, to perform the funerary act, or simply site, as known by archaeologists, will have to be analysed, once their components are physiographic and topographic elements. Within this natural category we must observe the visibility and intervisibility aspects trying to identify the places, or sites network, where they fit in, either of funerary or habitat nature. Other level of importance is expressed in the passage ways as in people and goods movements (Santos Estevez, Parcero Oibiña, Criado Boado, 1997: 63).

Van Gennep (1977) understands that in transition phases may have been a break in a previously social structure, followed by a waiting period and subsequently aggregation rituals towards a new structural line with incorporation of the new social status. According with Hertz (1960) there are several communities where death is not an instantaneous phenomenon. Instead, it conveys a transition process assisted with the proper ritual procedures. Turner (1967) sees that the decomposition of the flesh implies a kind of transition that the society itself has to go through. The death becomes, thus, after Huntington and Metcalf (1991) in a slow process allowing the community reorganization following the traumatic event that in some cases can cause a communitarian rupture considering, for example, the death of the leader.

We have, therefore, as established point that funerary rituals are a manifestation of a certain ideology in which the collective ideal as a firm position in social processes and, perhaps, political of the communities. To understand these ideological mechanisms Lewis Binford (1971: 19– 20) generated the concept of “social persona”, implying that there is two forms of understanding it. By one side, the ritual gestures and believes are the individual social identity representation, by other, the community elements acknowledge responsibilities in the dead person.

Observing the spatial relation between settlement and funerary area we observe in Souto case a segregation, being the tumulus isolated as Porto Escuro and Souto 5, or grouped with some few meters between, as in Souto tumulus 1 and 2, or grouped without spaces between the mounds as in Souto 3, 4 and 6 cases. Still the relations between tumuli itself, its landscape dispersion, the relation with the neighbouring settlements and with the soil use capacity, may be important to identify social identities and the choice of territory natural resources. The artefacts placed and now gathered inside the urn may be an indicative of social hierarchy or of goods exchange. The degree of information collected in the intervened sites can gives us a clue about the relation, on one hand of the caves landscape localization (all of them turned to Nabão river) so, to the water, and this element might have come determinant to the burial ritual; also in the megalithic monuments, with a great landscape location, with the possibility of being seen from far, mainly from the Zêzere river left margin.

The funerary modalities study and analysis makes us think that the connection to the ancestors may well have been used to legalize or negotiate power. Ian Hodder (1982) analysis three aspects related with funerary practices a) the neighbouring of cemeteries with the settlements, b) the depositions dispersal pattern, and c) the burial pattern. The Canteirões funerary rituals as, previously mentioned, may be divided into individual burials and in big areas of bones distribution, seem to obey an existing rule, not only to the individual burials but also to the collective ones. It is, however, obviously a change in the social status that initially gives priority to individualize death and that, soon after, still in the V millennium B.C.E. breaks through with a different ontological nature preparing the burial places towards collectivization, as given dominance to a place that for generations has been a burial place, either with as secondary burial for the bones or simply by placing a single body in fetal position. At the Bell-beaker period there is a change in the death

It is likely we are towards a situation in which the power comes from the social group in the way as Renfrew (1976) called chiefdoms grouped oriented. As suggested by Hodder (1982) is also possible that the three branches 70 

 

the individual treatment of the body we reflexively honour the person. We are obliged to reflect on the possibility of, when we see the body individually treated; to be facing a system where the propriety is important and related with inheritance transmission, in which society is controlled increasingly by parental ties.

mentioned by him are not as revealing of the represented social relations. The architecture is seen in a determined pattern of materializing the sites spatial organization. It reveals us a malleable interpretation once is assumed an interaction between the dominant Man and the dominated Space. In this aspect, the chronological time as also as ally the societies degree of technological development as well as the function type, either pragmatic or transcendental. At the same time the human activity expresses about ideology, over communication turning traditions and social relations conditioned to a chronological context. The architecture is then conceived as a human way of spatial integration and socialization and, then, also a materiality to account for in space and time contextualization. The space organized needs the society’s direct action over the environment and Nature revealing in each action or process a historical sense.

If we choose to consider the funerary monuments as memory keepers (collective or individual) we may be facing a situation as referred by Bradley (1993) in which there is a difference between the daily time and the ritual time. The ritual time will be a time existing in nonmeasurable abstract, imposing a causal nexus between the past and present becoming a transmitter of tradition and culture. In the majority of megalithic funerary chambers and in cave human bones were found. The bodies were laid and after natural decomposition, the human bones were reorganized spatially in the chamber making a real bone assets archive (Thomas, 1993: 35).

Several authors approach the Architecture and Space problem (Kent, 1990; Criado Boado, 1993a and 1993b; Tilley 1994; Bradley, 1998; Sanches, 2000; Valera, 2000; Jorge, et. al., 1998-1999) seeing the Architectural integration as humanized element that, might have a practical as well as a symbolic sense. Concerning the theme under analysis we must induce that funerary monuments implantation either in cave, in megalithic monuments and in tumuli, were conceptually forecast, having been assessed and selected the local, or site. In this way it has a landscape topographically reading, the visibility or intervisibility, even a geological substrate assessment, or the presence/absence of the water lines establishing thus a reading and communication of an element that might have been considered transcendental. It really seems to have had one preoccupation in the use of forms and raw materials as in the case of the megaliths construction.

The “decadence” and “death” of the megalithic phenomenon meaning must occur around the Bronze Age, towards the structures preservation or rebuilt. It is, however, to be noted, traces of votive and funerary occupation within the Bronze Age. We think that such occurrence is responsible for the loss of identity from the communities relatively to the megalithic cult places, turning it secondary, persisting, however, over the old ideological notions a new line of thought in death management. At the Megalithism climax we see that exists a sense or a message transmitted through monumental constructions; sense that can be revealed through territorial appropriation by the Neolithic agro-pastoral communities. If by one side, was an attitude of impose territoriality, we also see that their density and clear visibility transformed and invented a new landscape. It is highly probable that a new ideology (Final Bronze Age), when transforming the landscape, not only in the monumental aspect, as also made a reading of the physical space through the necessity of opening cultivation and grazing areas.

The death causes in human being a sense of vulnerability that, joined with the unknown fear causes a unique treatment towards the deceased relative. Parallel to this insecurity there is a will of organize in a systematic manner the place of the dead, either in ostensive mode, as in the case of the megalithic monuments, either in a landscape concealment, as in the case of the necropoliscaves and tumuli.

We think that Bronze Age communities, as Souto, have inherited a system of believes based in Neolithic ones and also a specific territory planning markedly Neolithic and Calcolithic. This inheritance was so strong that influenced the continuation, in some cases, of burials in the megaliths. It is likely that the sense of ownership was reflected in oral tradition and in symbolic associations with death. As observed, the burials in megalithic monuments are not compatible with the system of believes in the Bronze Age populations, yet, a new form of seeing the world comes without doubt not only with incineration but, and once more in Souto case, also with a new monument invisibility of the funerary space translating those new conceptions of the symbolic and of the form of interpreting death although architectonically the mound notion is maintained.

It is in a society defined by certain parameters that we observe the death treatment as a privileged individually management, although remaining the concept of collective death. That approach may have assumed forms of religiosity were the individual is increasingly important, and this attribute, is shown in the way as the body is handled. However, archaeography reveals us that the body was always important in the process, independently from the type of burial, collective, pit or ossuary. Looks as if, in Megalithism, during the Neolithic and Calcolithic, we revere the common ancestor hosting him in the chambers and corridors, whereas when we observe 71 

 

Thus, water, primarily element in species survival, where sapiens is included, it is, so we think, the primordial element of live, that transforms, shapes, not only the karstic cavities in the literal sense of the word but also, in figurative speech and, periodically, gives back live. There is a symbolic relation in the possibility of matching the functional site (cave-necropolis) with water (the Nabão river), that is to relate death in the mythological plan, donating the body to a life given place. This relation is quite intelligible, so, without a tight separation between the living world and the mythological universe of dead.

The megalithic monument case is not an exception; we will go, in other areas, to see the continuation of funerary use of the caves within this period. The case of the natural cavities becomes, on other hand, very interesting once it is related directly and in a first plan with our sensorial activity. Immediately, entering in these spaces, sometimes dark and silent, and often very humid, a psychological “preparation” is required to leave the light behind and enter inside, also it is need an emotional preparation to face the adaptation and orientation that will differ from that we are used to outside. It is, therefore, a thin line dividing the outer plan from the caves inner plan showing the existence of two different kinds of nature. The past contexts studies given by archaeography transmits us a perception of how the believes system and the symbolic universe was built generating one possible comprehension of the manner as death was seen in recent prehistory.

The water would have been necessarily related to the ritual gestures of the body deposition, gestures unfortunately lost for us but, that we can foresee some through the archaeological record, that is, a pit opened in the sediment to place a body, maybe in foetal position and its subsequent covering with the initial removed sediments.

We may deduce, before all, that a common participation from the community and close relates was present, so, we will be towards a participative and active attitude from all the group members in every stages of the event, that is, since the moment of death, including all the posterior moments, until all the rituals were completed.

If by one hand, it is possible to extrapolate minimally obviously conclusions related to man and its environment, and with water to a funerary imaginary world, relating it with colour is a problem considering that in terms of field data wasn’t found a human bone painted with ochre. However, we have recovered some ochre nodules with red pigmentation in the Gruta do Cadaval C layer, in the first and second room.

In Souto the attention provided to the body instils a particular bond towards the family and community in a way that the ritual process sets a new postmortem solidary stage, until the moment in which the body is recycled trough fire. Perhaps was for it that the “places of dead” were raised, sometimes nearby the community, sometimes in specific places apart from the living environment (caves, dolmens, tumuli), but present in the social and symbolic memory of the group. The umbilical cord that unites the living community to that deceased persona is highlight trough processed mourning rituals, we think, by the closest family and by all community members tied with him by brotherhood.

The presence of these nodules just by itself can give us another dimension of this raw material use as supporting element of the votive elements collected in these collective burials layer. So, in Canteirões, there is a primordial and direct relation of the burials with water, by one hand and with colour, by other. Following to East venturing into the megalithic world we have our monuments placed in the lower Zêzere valley. This area is formed by ridges relatively high not over 400 meters (higher than limestone area) of small mountain type, with very steep abrupt slopes, going directly towards the river. In it we find in the Zêzere right margin Dolmen 1 de Val da Laje and Dolmen das Pedras Negras and in the left margin the Alqueidão orthostats, Vale Chãos necropolis, Jogada necropolis, Pedra da Encavalada and Medroa menhir.

Towards conclusion, or maybe not? Finally and, beyond the remarks over field data and symbolic characters that we have made, we would like to relate the natural carved structures with the ones built by Man in its environment, and also with the colour. Now let us approach the relation that we may establish between caves-necropolis and its environment. In this frame, we may observe sedimentary formations compact, fissured, with variable permeability, in rift areas proximity, belonging to the alpine orogeny cycle. The soils are classified as lithosols associated with luvisoils; they are soils mainly alkaline in the F agricultural usage scale, in a phytoclimate Atlantic Mediterranean area, characteristics common to the entire Canteirões area.

Here Man’s relation with death is made, not through symbioses with sculpted nature, but through massive construction using gneiss bedrocks abundantly present in the area. We know that this region has porous, very low permeability formations, with schist-greywacke sedimentary rocks and metamorphic series geologically formed in Pre-Cambrian era. The soils classification is orthic podzols in association with eutric cambisoils, acid and from F class, being the natural region globally a basal phytoclimate one.

Beyond this global physiographic analysis, common and scientific, another strand must be studied and it is directly connected with the funerary imaginary and with the fact that all the cavities entrances are faced to Nabão River.

These physiographic characteristics give us one possible tendency to its use as forest within the megalithic main 72 

 

area in a phase where, we think, the cattle and some incipient agricultural would have been already expanded throughout the valley.

The relation is naturally established by the body recycling in ashes that later are transferred to ceramic urns. This new ritual gestures procedure, maybe even more complex, once also the community becomes more complex adding to the agro-pastoral economy the metallurgic work with fire and using it as purifying element by itself.

The relation between stone receptacles construction with a morphology similar to an habitat, with a mound made of sediment and a stone ring, makes us rethink this new style of considering and elaborate death with roots in other natural elements, also primordial to the agropastoral communities, the land. So, the megalithic death of the lower Zêzere valley would be inevitably rendered to the earth element from where live that feeds the human beings is extracted. Once more we find death strictly connected to life through the reborn and cycle renovations of the seasons. Also here we see a direct relation to a natural element that allows the existence of several life forms, being death and life included in the same cycle.

We may then, as end, to understand the bi univocally, cyclical and inescapable relation between the living “world” and the dead “world” adding the connotative elements present in nature (water, earth and fire) and supporting at an imaginary level all these human communities mythology. Thus the Alto Ribatejo men in recent prehistory raised myths, histories, legends where the intrinsic relation between Man and Nature stays alive during millenniums.

Walking further northeast of this geographical area we come across with Souto necropolis, which is not apart. In topographic terms the small mountain chains geomorphologic modelling is maintained and precisely in the small plateau we find these monuments discrete relatively to the landscape visibility and also built by Man.

Those myths and legends are inaccessible to us; even when grammatical study of rock art or burial patterns can give indict evidence of its presence. Although, is possible to record in all this process, marked by differences that cross socio-economical very distinct contexts (from early Neolithic to final Bronze), interesting convergences and continuities that, far from any environmental determinism, seems to express the relevance of the territory in the human behaviour organization.

At this point was chosen, by final Bronze Age communities, a new element. We are not only facing the evidence of the earth element, to this is added another element, the fire. the fire.

References ALLUÉ, E. (2000) – Pollen and charcoal analyses from archaeological sites from the Alto Ribatejo (Portugal). In CRUZ, A. R.; OOSTERBEEK, L., ed. – Arkeos, Territórios da Pré-História em Portugal. Tomar: CEIPHAR, nº 9, p. 37-57. ALMEIDA, N. (2010) – A Constituição das Primeiras Economias Agro-Pastoris, Paradigmas em Debate: o contributo da Zooarqueologia e Tafonomia para o Alto-Ribatejo. Dissertação de Mestrado inédita. Universidade de Tras-os-Montes e Alto Douro. 291 p. BINFORD, L. (1971) – Mortuary Practices: Their Study and Their Potential. In BROWN, J. A. (ed.) – Approaches to the Social Dimensions of Mortuary Practices. Washington DC: Memoirs of the Society for American Archaeology, nº 25, p. 6–29. BRADLEY, R. (1993) – Altering the Earth. The origins of monuments in Britain and Continental Europe. Edimburgo: Society of Antiquaries of Scotland. BRADLEY, R. (1998) – The significance of the monuments. Londres: Routledge. CERRILLO CUENCA, E. (1990) – Arqueología de las Religiones Primitivas y Arqueología de las Religiones Organizadas. Una Reflexión. Zephyrus. Salamanca: Universidad de Salamanca, nº 43, p. 189192.

CRIADO BOADO, F. (1993a) – Visibilidad Interpretación del registro arqueológico. Trabajos de Prehistoria. Madrid: Consejo Superior de Investigaciones Científicas, nº 50, p. 39-56. CRIADO BOADO, F. (1993b) – Límites y posibilidades del paisaje. SPAL. Revista de prehistoria y arqueologia. Sevilla: Universidad de Sevilla, nº 2, p. 9-55. CRUZ, A.; OOSTERBEEK, L. (1985) – Gruta do Cadaval - 1983. Informação arqueológica. Tomar: Câmara Municipal, volume 5, p. 117-8. CRUZ, A. (1997) – Vale do Nabão. Do Neolítico à Idade do Bronze. Tomar: Centro Europeu de Investigação da Pré-História do Alto Ribatejo, Arkeos, perspectivas em diálogo, volume 3, 361 p. CRUZ, A. (2000) – Necrópoles de Gruta no Contexto da Neolitização do Alto Ribatejo. In SANCHES, M.; ARIAS, P. (coord.) – Neolitização e Megalitismo da Península Ibérica. Porto: ADECAP, Actas do 3º Congresso de Arqueologia Peninsular, III, p. 61-79. CRUZ, A. (2008-2009) – Relatório da campanha de escavação da Mamoa ou Tumulus 1 do Souto – 2008. Lisboa: Portal do Arqueólogo. CRUZ, A. (2011) – A Pré-História Recente do vale do baixo Zêzere. Tomar: Centro Europeu de Investigação da Pré-História do Alto Ribatejo, 73 

 

Gruta do Cadaval (Tomar). Actas da 1ª Reunião do Quaternário Ibérico. Lisboa: Instituto Nacional de Investigação Científica, volume II. OOSTERBEEK, L. (1985b) – Elementos para o estudo da Estratigrafia da Gruta do Cadaval (Tomar). Almadan. Almada: Centro de Arqueologia de Almada, volume 4/5, p. 7-12. OOSTERBEEK, L. (1987a) – Gruta do Cadaval. Informação Arqueológica. Lisboa: IPPC, volume 8, p. 79-80. OOSTERBEEK, L. (1987b) – Gruta dos Ossos. Informação Arqueológica. Lisboa: IPPC, volume 8, p. 80-81. OOSTERBEEK, L. (1987c) – Projecto de Estudo da Neolitização do Vale do Nabão. A Gestão dos Espaços e os Métodos de Abordagem. In MANIQUE, A. P. – Temas de História do Distrito de Santarém. Santarém: Escola Superior de Educação de Santarém. OOSTERBEEK, L.; CRUZ., A. (1990) – Gruta do Cadaval - Gruta dos Ossos - Anta 1 de Val da Laje. In: VELOSO, C.; PONTE, S. (coord.) – Imagens de Tomar - Roteiro Histórico. Tomar: Secretariado do VIII Encontro de Professores de História da Zona Centro, p. 15-20. OOSTERBEEK, L.; CRUZ, A. (1991) – A Arqueologia da Morte: considerações a propósito da interpretação dos contextos sepulcrais na região de Tomar. Boletim Cultural da Câmara Municipal de Tomar. Tomar: Câmara Municipal, volume 15, p. 267-291. OOSTERBEEK, L. (1991) – Carta Arqueológica de Tomar. Revista de Ciências Históricas da Universidade Portucalense Infante D. Henrique. Porto: Universidade Portucalense Infante D. Henrique, volume VI, p. 77-89. OOSTERBEEK, L.; CRUZ, A.; FELIX, P. (1992) – Anta 1 de Val da Laje: notícia de 3 anos de escavações (1989-91). Boletim Cultural da Câmara Municipal de Tomar. Tomar: Câmara Municipal de Tomar, volume 16, p. 31-49. OOSTERBEEK, L. (1993a) – Gruta dos Ossos (Tomar). Um ossuário do Neolítico final. Boletim Cultural da Câmara Municipal de Tomar. Tomar: Câmara Municipal de Tomar, volume 18. OOSTERBEEK, L. (1993b) – Nossa Senhora das Lapas – excavation of Prehistoric cave burials in Central Portugal. Papers of the Institute of Archaeology. London: University college London, volume 4, p. 4962. OOSTERBEEK, L. (1997) – Echoes from the East: late prehistory of the North Ribatejo. Tomar: CEIPHAR, Arkeos, perspectivas em diálogo, volume 2, 304 p. RENFREW, C. (1976) – Megaliths, Territories and Populations. In DELAET, S. (ed.) – Aculturation and Continuity in Atlantis Europe. Bugres: de Tempel, p. 198-220. SANCHES, M. (2000) – As gerações, a memória e a territorialização em Trás-os- Montes (Vº - Iº mil. AC). Uma primeira aproximação ao problema. PréHistória Recente na Península Ibérica. Porto: ADECAP, Actas do 3º Congresso de Arqueologia Peninsular – 1999, p. 123-146.

Arkeos, perspectivas em diálogo, volume 30, dissertação para obtenção do grau de Doutor, Universidade de Trás-os-Montes e Alto Douro, 273 p. CRUZ, A.; GRAÇA, A.; OOSTERBEEK, L.; ALMEIDA, F.; DELFINO, D. (2013) – Gruta do Morgado Superior – Um Estudo de Caso Funerário no Alto Ribatejo (Tomar, Portugal). Las Crisis en La Historia: Noción y Realidades. Comunidad de Castilla-La Mancha: Departamento de la Universidad de Castilla – La Mancha, Vinculos de Historia, nº 2, URL: vinculosdehistoria.com/índex.php/vínculos/article/vi ew/62, p. 143-168. CRUZ, A; OOSTERBEEK, L. (1998) – Anta 1 da Jogada (Abrantes). Techne. Tomar: Arqueojovem, volume 4, p. 26-31. CRUZ, A.; OOSTERBEEK, L. (1998) – Anta 2 da Jogada (Abrantes). Techne. Tomar: Arqueojovem, volume 4, p. 32-35. CRUZ, A.; OOSTERBEEK, L. (1998) – Anta 3 da Jogada (Abrantes). Techne. Tomar: Arqueojovem, volume 4, p. 32-35. CRUZ, A.; OOSTERBEEK, L. (1998) – Anta 4 da Jogada (Abrantes). Techne. Tomar: Arqueojovem, volume 4, p. 36-41. CRUZ, A; OOSTERBEEK; L. (1998) – Anta 5 da Jogada (Abrantes). Techne. Tomar: Arqueojovem, volume 4, p. 42-48. DREWETT, P.; OOSTERBEEK, L.; CRUZ, A.; FÉLIX, P. (1992) – Anta 1 de Val da Laje 1989/90 - The excavation of a passage grave at Tomar (Portugal). London: Bulletin of the Institute of Archaeology. FÉLIX, P. (1993) – A Região Nabantina no Final da PréHistória. Boletim Cultural. Tomar: Câmara Municipal, Boletim Cultural da Câmara Municipal de Tomar, nº 19, p. 237-254. FERREIRA, A. (n.d.) – Caracterização de Portugal Continental. (Disponível em url: http://repositorio.lneg.pt/bitstream/10400.9/542/2/ca p2.pdf. Consultado em 12.04.2013). HERTZ, R. (1960) – Death and the Right Hand. Londres: Cohen & West. HODDER, I. (1982) – Symbols in Action: Ethnoarchaeological Studies of Material Culture. Cambridge: Cambridge University Press. HUNTINGTON, R.; METCALF, P. (1991) – Celebrations of Death: the Anthropology of Mortuary Ritual. Cambridge: Cambridge University Press. JORGE, S.; OLIVEIRA, M.; NUNES, S.; GOMES, S. (1998-1999) – Uma Estrutura Ritual com ossos humanos no sítio pré-histórico de Castelo Velho de Freixo de Numão (Vª. Nª: de Foz Côa). Portugália. Porto: Universidade do Porto. Faculdade de Letras, Nova Série, volume XIX-XX, p. 29-70. KENT, S. (1990) – A cross-cultural study of segmentation, architecture, and the use of space. In KENT, S. (ed.) – Domestic Architecture and the Use of Space: An Interdisciplinary Cross-Cultural Study. Cambridge: Cambridge University Press, p. 127-153. OOSTERBEEK, L. (1985a) – A Facies Megalítica da 74 

 

SANTOS ESTEVEZ, M.; PARCERO OUBIÑA, C; CRIADO BOADO, F. (1997) – De la Arqueología Simbólica del Paisaje a la Arqueología de los Paisages Sagrados. Trabajos de Prehistoria. Madrid: Consejo Superior de Investigaciones Científicas, volume 54, nº 2, p. 61-80. TILLEY, C. (1994) – A Phenomenology of Landscape places, paths and monuments. Oxford: Providence. THOMAS, L. (1993) - Antropología de la muerte [1979].

México: F.C.E. VALERA, A. (2000) – Em torno de alguns fundamentos e potencialidades da Arqueologia da Paisagem. Lisboa: Era/Colibri, Era Arqueologia, n°1, p.112121. VAN GENNEP, A. (1977) - Rites of passage. London: Routledge and Kegan Paul.

75 

 

76 

 

Collective Burial Caves in Spanish Extremadura: Chronology, Landscapes and Identities Enrique Cerrillo Cuenca, Antonio González Cordero

phenomena to the building of megalithic graves (Barnatt & Edmonds, 2002; Beyneix, 2012; Boaventura, 2011; Cerrillo-Cuenca & González, 2007). A key question is if the communities that used both kinds of funerary modalities shared a common cultural base and a similar conception of social relationships. Although, the analysis of interred individuals could shed light on a more accurate definition of the problem, the reality is that the available sources for comparing populations from caves and megalithic burials are scarce for Extremadura region (Figure 7.1), where this paper focuses. This last assumption would probably explain why archaeologists from not only Extremadura, but the surrounding regions, have dealt with items from material culture rather than with anthropological data. Focusing on the most elementary term of comparison, the grave goods, it soon becomes clear that no differences can be outlined between the artefacts deposited in natural caves and those from megalithic graves. Regarding this, Portuguese archaeologists, like V. Gongalves (1978), have employed the expression megalitismo de grutas (“megalithism” in caves) for referring to the fact that, in their most formal aspects (i.e. material culture), there is not an apparent distinction between the grave goods from either type of sepulchres.

Abstract In this paper we present the recent advances on the characterization of burial caves from Extremadura region (Spain). We present an appraisal of actualised data, which includes recent absolute dates and anthropological data. We briefly discuss about the interpretation of collective burial caves and the observed rituals from Early Neolithic to Copper Age. Key-words: Burial Caves, Neolithic, Copper Age, Collective Burials 1. Introduction Collective burial caves can be found in almost every location of Prehistoric Europe, representing one of the most common forms of burials in the Atlantic countries. The archaeological works developed in territories like Ireland (Dowd, 2008), Britain (Chamberlain, 1996), France (Beyneix, 2012), Portugal (Araujo et. al., 1995, Cruz, 1997, Carvalho, 2007, Tomé & Oosterbeek, 2011; Zilhão, 1992, among many others) and of course, Atlantic and Mediterranean Spain, have revealed that the use of natural burial places of burial represent a synchronic

Figure 7.1: Location of Extremadura region in Iberian Peninsula and archaeological sites cited in the text. 77 

 

funerary use of the caves during Prehistory. For example, sites as Boquique (Plasencia, Cáceres), probably one of the most famous rock-shelters from the Neolithic period of inner Iberia, might have acted as funerary containers during the Copper Age, especially if we consider some of the artefacts that were recorded in a paper presented by P. Bosch-Gimpera (1915-1920) almost a century ago. Adzes and bowls in serpentine rock, among other features, can be seen as representative from this possible funerary function, along with the remains of non-identified cremated bones (García, 1915-1920). Although, such an attribution can be speculative, the differences in the preservation of osteological evidences should be taken into account when tracing the distribution map of burial caves. In this regard, the distribution of this kind of funerary context in the Extremadura region is formed by almost 20 caves or shelters, where the funerary use of caves can be assured with some degree of certainty.

In the Iberian Peninsula, scholars have often seen burial caves and megalithic sites as purely independent contexts, generating alternative discourses for explaining a not so apparent duality. This duality has often been perceived in terms of chronology, suggesting that the use of burial caves would have preceded the building of passage graves and it did not last much longer. Such a statement can be found in the some works, when in fact the available chronologies suggest that the funerary use of caves persisted until the final moments of Late Prehistory, if not beyond that moment (Esparza, 1990; Hierro, 2011). Other scholars have presented the duality as the product of different capacities for labour force mobilisation, or even as a sign of the segmentary condition that Neolithic communities face to more complex ones during the Copper Age. Finally, some authors have seen in this polarity the prevalence of two different systems for land exploitation and resource management (Tomé & Oosterbeek, 2011). In this regard, multiple kinds of interpretations can be addressed, even in functional terms (e.g. access to the adequate geological background for obtaining building materials). The access to identitary matters might be complex in the current state of art, and the question can perhaps be addressed through the treatment of corpses and the disposition of grave goods. This can be in fact a more thought-provoking research line, although the available contextualised data is fragmentary and scarce. Although, megalithic monuments are still today the most prominent kind of burial, the recovery of bones was not possible up to the date, so the closer references have to be found in the nearby regions of Portuguese Alentejo and Beira Alta, where extensive reports and references have been published along the years (Tomé, 2010). In this regard, the presentation of osteological unpublished data from Tío Republicano serves here as a new reference for comprehending the nature and ritual of collective Copper Age burials in Extremadura region.

Historiographical data can contribute to a better presentation of data regarding the disposal of bones into cavities and shelters. Some textual sources give account for the disposal of individuals on the ground of cavities with no further treatment except the sealing of sepulchres with walls made of mud and stone (Cerrillo-Cuenca & González, 2007). However, some of these features have not have been confirmed by modern archaeological fieldworks. Collective burials, as we will show below, are in any case the most recurrent way for disposing the individuals. In fact, in the inner areas of the Tagus Basin single burial have been documented in provinces like Madrid (Martínez, 1984) or Guadalajara (Jordá & Mestres, 1999), which gives credit to the idea of widespread phenomena. Ossuaries are perhaps the most well documented context, and were reported in several cases, for instance, in Maltravieso cave, dated to the Bronze Age (Cerrillo-Cuenca et. al., 2008). 2.2. The Archaeological Evidence

2. Spanish Extremadura: An Appraisal of Evidence In Spanish Extremadura a new absolute date for burial caves comes from Cueva de Postes (Fuentes de León, Badajoz), where a date from the beginnings of 6th millennium cal BC has been briefly presented (Collado & García, 2011). The burials were practiced in limestone caves, whose sequence has been superficially published. Regarding the context, most of the data is still unknown and no further information about the burial is provided. Nevertheless, such data is important so as to acknowledge the existence of Mesolithic settlements in inner areas of the Iberian Peninsula, which increases the number of sites already published by us some time ago (Arias et. al., 2009a). Thereby, Mesolithic cave burials seem to be a more widespread phenomenon than it was thought, being only recognisable some decades ago in the Northern fringe of Spain or the Levantine area (Arias et. al. 2009b). Its recording at the inner areas, as the burials from BrañaArintero (Vidal et. al. 2010) can prove, and in the South of Spain, where the data from Nerja (Jordá & Aura 2008) stand as the solely reference gives support to the idea of the preference for natural places for inhumation. Thus, the data from Cueva de Postes can shed light on the

2.1. The Historiographical and Previous Evidence The archaeological record from the burials caves known in Spanish Extremadura is extremely biased, and it should be noted that until the 1950s we have account for burial caves that were not documented through an archaeological methodology. Historiographical information (Cerrillo-Cuenca & González, 2007) reveals the great amount of information that was cited in this subject since the 16th century, which is no longer accessible. In this sense, burial caves seem to be far more widespread than what is recorded today from an archaeological point of view, but pillaging and explorations from early antiquarians made a significant part of cases disappear. Besides, post-depositional factors have affected the preservation of human remains at rockshelters from certain geological environments where the acidity of soils does not favour the preservation of bones, like granite rock-shelters (Calado, 2000), or fragile stratigraphic contexts on schist or quartzite. Because of this certain artefacts (beads, adzes, etc) might prove the 78 

 

Neolithic burials in inner Tagus basin. Our works in the Canaleja Gorge have provided additional data about this funerary behaviour during the last decade, amplifying the possibility of recognising the continuation of burial caves during the earlier stages of neolithization. The cave complex in Canaleja Gorge is formed by at least five caves and small shelters located at a narrow fringe of limestones (Algaba et. al., 2000), three of which have been excavated by us. The caves are elevated a few metres over a tributary stream of the Tagus river, named Canaleja, and are located in the bottom of the skirts of a steep valley towards which their entrances open. In this case, the excavation of Canaleja 2 (Figure 7.2) site yielded bone remains from three individuals without anatomical connection (Figure 7.3). A well-preserved jaw and bones from a foot belonging to an adult male, and three vertebrae from a juvenile and a remains from a leg from an infantile were recovered (Barca, 2007). The reduced number of bones from each individual suggests that the place was used as a primary burial site, which was later opened to recover the skeletal remains. This funerary use might be compatible with the closing wall that was documented at the entrance of the shelter.

origins of a funerary custom that is unprecedented in Extremadura, but also in areas with a significant amount of burial caves like West-Central Portugal, where the archaeological record of cave burials during Late Prehistory is far more enough expressive than that of Extremadura. One interesting question is how these occupations are correlated with the data from the Southern side of Sierra Morena mountains, where Pellicer and Acosta (1982) excavated the caves of Santiago de Cazalla, obtaining similar absolute dates for layers erroneously interpreted as belonging to the Early Neolithic. In the Tagus basin, the existence of the Mesolithic has been outlined during the last years, thanks to the discovery of Early Mesolithic (or Epipaleolithic) sites, although no funerary evidences have been recorded at the documented sites. Early Neolithic cave burials have also been documented in the Iberian Peninsula, although in a low number. The data from Can Sadurní (Blasco et. al., 2005) in Catalonia is perhaps amongst the most complete data, although not the only (García et. al., 2011). The Nossa Senhora das Lapas site, whose works were published by Oosterbeek (1993) twenty years ago, was the only reference for Early

Figure 7.2: Plan of Canaleja 2 site.

79 

 

Figure 7.3: Osteological remains recovered at Canaleja 2. during the Early Neolithic era, as impressed pottery was found during the excavation. However artifacts from the Copper Age and the Bronze age were also recovered, which would prove that the cave might have served as a funerary container in later chronologies. In order to obtain an approximate date of the burials a parietal from an infantile individual was dated, which rendered a dating from the beginnings of 4th millennium cal BC (Beta202343, 5100±50 BP, 3980-3780 cal BC) a moment that is considered in the regional sequence as Middle Neolithic (Cerrillo-Cuenca, 2005). No diagnostic artefacts were collected for this period, except for microliths, whose relative dating can be considered ambiguous. The data is excessively fragmentary, although it can be inferred from them that the funerary caves in Spanish Extremadura are coeval to the very first megalithic monuments of the Tagus basin, whose chronologies point to an early development around the transition from 5th to 4th millennium cal BC (Bueno, et. al., 2005; Ruiz-Gálvez, 2000).

A paleo-pathological interpretation of these remains (Barca, 2007) has been shortly published, although the scarcity of the collection does not allow an in-depth analysis of pathologies. Isotopic analyses were performed on a tooth from the adult individual, showing a mixture between marine and terrestrial signals (d13C: -17,7, d15N: 9,46). Although the data might fit into the published intervals for Neolithic populations (Lubell, et. al., 1994), the data stands as an isolated sample with a certainly poor statistical representation for the Neolithic communities of the inner basin of Tagus. Moreover, further isotopic analysis from faunal bone remains would be needed to discard alternative environmental factors that could have influenced isotopic relations. Paleo-environmental data from the Canaleja 2 site are quite similar to those provided by Early Neolithic sites from Spanish Extremadura (Cerrillo-Cuenca, et. al., 2010), especially regarding human intervention on the environment and the development of agricultural practices. Although the intensity of cropping practices is unknown, it seems feasible that a low-intensity activity was carried out in the surroundings of the shelter. These evidences of cereal croppings, during the Early Neolithic, are compatible with a very possible domestic use of space attested by fragmented impressed pottery (Boquique type) and handstones for saddle querns. An absolute dating from dispersed charcoals offered the following date: AA78257 6203±44 BP, (5300-5043 cal BC).

At Tío Republicano Cave (Figure 7.4, 7.5), the analysis of bone remains has shown a more complex inhumation ritual, where cremation is involved. Disarticulated and fragmented bones were found forming an ossuary in the main room of the cave with no apparent pattern in its disposition. A thin layer of clay covered the ossuary. The most complete bones were the remains of two skulls that laid on the ground covered by a pile of small broken bones. No larger bones were recovered, from which it can be deducted that the ritual implied a conscious reduction of the size of the bones. It is worth noticing the inclusion of faunal remains in the sample, which can be due to either a deliberate inclusion or to post-depositional activity. The minimal number of individuals was established as 11, one of which is a juvenile and the other an infantile.

Canaleja 1 is located few meters apart from the aforementioned shelter. The cave is composed of a natural corridor that finishes in a narrow room. The deficient preservation of its stratigraphy makes its interpretation as a burial cave difficult, although human remains were recovered from within the cave. Many of the recovered artifacts suggest the occupation of the cave

80 

 

Figure 7.4: Location of Tío Republicano site at Canaleja Gorge.

Figure 7.5: Section and Plan of Tío Republicano Cave. bones or if the heating was performed when the bones were already excarnated. No additional evidence for cuttings on the surface of the bones was detected for assuring that the goal of the transformation was the defleshing of corpses. Secondly, no differences in the treatment between the types of bones can be observed, which can suggest that the intention was to achieve a general reduction of the individuals to bone remains. In this regard, is it almost impossible to identify the formation process of the ossuary, or if the cremation of bones may have been executed long after the individuals have deceased. As the anthropological reports states, it is

In all the cases that the remains had been exposed to a heating source, the different colours of their surfaces allow us to draw some conclusions (González & Rascón, unpublished). Extremely white bones indicate the high temperature reached during the cremation process. The temperatures, determined by comparisons with experimental tables (Etxeberria & Delibes, 2002), ranged from 200 to 600 degrees Celsius, reaching the point before incineration occurs. Several appointments can be made in this regard. Firstly, as the anthropological analysis have shown that it is not possible to establish if the intentionality of cremation was the defleshing of 81 

 

not impossible to assure if the bones were cremated once they were dried (González & Rascón, unpublished). The evidences for longitudinal splitting of bones, which are due to the exposition of dry bone to fire (Ubelaker, 2009), could not be clearly distinguished in the collected sample. Regarding the absolute dating of human remains, the laboratory failed to extract collagen due to the heating of the bones. Two different samples with no apparent signs of heating or burning were sent for dating lack of collagen, which implied that all the bones were exposed to heating. Thus, the approach to the chronology of the burial had to be done through the typological analysis of the grave goods. Large blades of flint (Cerrillo-Cuenca, 2009), flint arrow points, a bead made on variscite and an anthropomorphic idol cut from a rib bone (Fugure 7.6), are the most representative artefacts, which suggests a funerary use of the space during the 3rd millennium cal BC. The accompanying grave goods seemed not to be affected by the heating of fire, showing a differential treatment on their deposition. Flint elements do not present any kind of thermal extraction or signs of heating, like the rest of the elements deposited besides the osteological remains. Since no remains of fire or charcoals were detected at the cave, the manipulation of bones seems to have been carried out partly outside the funerary space.

Figure 7.6: Bone idol fromTío Republicano.

In spite of the fact that the cremation of human remains has been reported in some funerary contexts from Extremadura region, some references remain unpublished. As far as we know, cremations are documented in megalithic sepulchres like Lácara site, which gives credit to the information reported by the Leisners at the near Portuguese monuments of Olival da Pega and Anta de Cebolino (Leisner & Leisner, 1951). In the last years we have had the opportunity to document fragmented and burned human remains at Cueva del Valle site (Cerrillo-Cuenca et. al., in press). The cremated bones (Figure 7.8) from Cueva del Valle (Figure 7.7) and Tïo Republicano share a common pattern in their treatment, so the funerary custom would have been spread through a wide territory and not only circumscribed to the the basins of Tagus or Guadiana River. This rock-shelter consists of one room that was arranged as a sanctuary during Roman times. This latter use had dismantled the remains of the funerary occupation, depositing the remains at the entrance of the cave. The particularity in this case is that part of the grave goods seems to have been exposed to fire, as the thermal extractions show in the most part of the surface from the flint blades (Figure 7.9). The role that these specific tools might have played during the funerary ritual cannot be explained, as it is even difficult to establish if they were beside the corpses (or bone remains) during their cremation. However, flint arrow points seem to have been included among the bones with no signs of exposition to fire, revealing a different pattern of deposition.

In general terms, osteological remains from 3rd millennium populations are not represented by a significant amount of archaeological context in the region, except for the corbel-roofed monuments from the Guadiana River basin, like La Pijotilla (Hurtado, et. al., 2000) o Huerta Montero (Blasco & Ortiz, 1991), or the unpublished data from Cerro de las Baterías (Badajoz). Therefore, our knowledge might be circumscribed to a concrete type of burial monument, chronology and, consequently, to similar ideological and social parameters. Additional references come from the Guadiana River drainage, concretely from Charneca cave (Enríquez, 1986), a small rock-shelter opened on a quartzite wall. As in many other cases, the mixed stratigraphy did not allow its excavator to recognise the original disposition of the bones, nor the number of individuals being interred at the site. The recovery of an idol made on a large bone, whose specie is unidentified (Enríquez & Rodríguez, 1990), can be dated in the transition from Late Neolithic to Copper Age. Several sites can also be included in a wide temporal fringe that covers from the Early Neolithic to Copper Age, although their precise chronology is difficult to determine. Caves like El Conejar, known since the early works carried out in the beginnings of 20th century (Del Pan, 1917), have yielded human bones, despite that the stratigraphic context cannot be reconstructed. As it happened during the excavation at La Chanerca site, fragments from “schist plate idols” were also recovered. This fact relates the funerary use of both sites with the megalithic burial chambers from both the Guadiana and Tagus drainage, where these kinds of artefacts become a common characteristic since the Late Neolithic times.

The recognition of other burial caves used during the Copper Age in Spanish Extremadura is however scarce.

82 

 

Figure 7.7: Aerial Photograph (Photo: S. Celestino) of Cueva del Valle, indicating the place where the test pits were practiced.

Figure 7.8: Cremated human remains from Cueva del Valle. A) Sample of bones, B) Different colorations that evidence the exposure of bones to different temperatures, C) transversal views from bones showing different colorations in their inner parts. 83 

 

Figure 7.9: Fragments form flint blades recovered at Cueva del Valle site. Thermal extractions can be observed in almost all pieces. Finally, a case illustrating the usage of burial caves during the Bronze Age has been documented in Maltravieso cave. The information about the disposal of bone remains at Maltravieso is due to the oral information recorded in the moment of its discovery (Callejo, 1958). It seems feasible to admit that three ossuaries were located besides the wall of the cave with no order, forming small piles. The bones were not systematically gathered, so the number of the individuals that formed the ossuaries and their specific location remain unclear. A population of at least 11 individuals has been inferred (Muñoz & Canals, 2008) after the 2002 field season, which is roughly consistent with the kind of data that other burial caves have provided in previous periods.

Particularly interesting is the presence of a trepanned skull (Álvarez, 1985) (Figure 7.10), a feature that is still unique in this region. In recent years we have suggested that the first rooms of the cave might have been used as a collective burial during the “Proto-Cogotas I” period (Cerrillo-Cuenca & González, 2007; Cerrillo-Cuenca, et. al., 2008), that is, during the central centuries of 2nd millennium cal BC. This dating is based on the shapes and motifs of decorated pottery, which clearly belongs to this period, but also on the inclusion of a bronze lance head among the recovered artefacts. However, recent works have claimed that the ossuaries are dated before Bronze Age (Muñoz & Canals, 2008), an opinion that is not based on any archaeological argument.

Figure 7.10: Trepanned skull (posterior view) from Maltravieso Cave (Photo. B. Sánchez). 84 

 

obtention of material for the building of dolmens, as is documented by the cases of Magacela (Bueno & Piñón, 1985). A similar situation is described for the surrounding area of El Conejar and Maltravieso caves in Cáceres city, where a series of megalithic barrows is known in the proximities of the city (Cerrillo-Cuenca, 2008). The same coincidence occurs at Chiquita Cave (Cerrillo-Cuenca & González, 2007; García, et. al., 2011), where shelters with rock-art paintings and evidences for inhumation have been recorded, in an area where megalithic barrows are well-known in the narrowing area (Bueno, et. al., 2011).

3. Discussion Taking into consideration the presented data, it seems evident that natural caves have been used as funerary containers for a long span of time, from the Mesolithic to the Bronze Age, probably suffering variations in their meaning and the social behaviour related to their use. The assimilation of megalithic phenomena might have occurred during the transition to 5th from 4th millennium cal BC, as the absolute dates from Canaleja 2 and Cueva de Postes suggest. From a material point of view, the relationship between burial caves and megalithic sites is undeniable. The most recurrent element in funerary contexts such, as schist plates, are equally present in burial caves like El Conejar or La Charneca, and in some cases in cut-shaped bones, like in Tío Republicano.

Regarding paintings, we suggested in a recent paper (Cerrillo-Cuenca & González, 2011) that the shelters decorated with schematic art could have served as funerary sites. Indirect information comes from caves where human bones were registered in association with rock-art (Cerrillo-Cuenca & González, 2007), although we still lack the stratigraphic contexts that could confirm a direct relationship between the graphical evidence and their use as burials. No many decorated shelters from Extremadura have been excavated up to the present date. An exception could be Arroyo Estanque shelter (CerrilloCuenca, 2011), where the excavation rendered the remains of a Neolithic habitat (Cerrillo-Cuenca, et. al., in press). More concrete evidence comes from the basin of the Guadiana River, at La Charneca. A few meters away from the entrance of the cave, a set of “anthropomorphic schist plates” are depicted on the rock (Collado, et. al., 1997), reinforcing perhaps the funerary meaning of the place (Collado, et. al., 1997: 148). The evident connection between the painted motifs and the funerary place, where schist plates were present (Enríquez, 1986), shows how schematic paintings might be involved in defining and characterising the burial space.

Regarding their insertion in landscape, burial caves and megalithic sites seem to share the same territories in a broad sense. In fact, this question is far from being clarified with simple arguments, since hybrid funerary solutions do exist in Spanish Extremadura, where this paper is focused. For example we have information from a site in the northern side of Sierra Morena, where a megalithic chamber was reproduced by disposing schist slabs forming a chamber in the inner part of a sediment voided outcrop (Prada & Cerrillo-Cuenca, 2006). The “scenification” of a megalithic chamber into a natural space also recalls the importance that the natural place itself holds, just in the sense that Scarre (2009) has proposed for megalithic sites. In this regard, several authors have recalled that in its formal architectonical aspect a megalithic monument resembles and reproduces itself as the concept of a cave. A brief analysis would provide more functional features that could be outlined, like the walls that enclosed the entrance to monuments and caves (Cerrillo-Cuenca & González, 2007). Moreover, a direct link between caves and megaliths could have existed, although the records that we have do not let us make such kind of inferences for most occasions. We could cite here cases in the Atlantic world where the bone remains of individuals have been transported from caves to megalithic chambers (Barnatt & Edmonds, 2002: 114), being perhaps a more common behaviour than we can actually identify in the archaeological record of monuments in Iberia.

To sum up, although the natural features of the terrain must be considered in a local scale, the proximity between both kinds of burials makes it difficult to defend substantial differences in terms of culture or subsistence strategies, at least for Spanish Extremadura. Unfortunately, paleoenvironmental data is not available for the majority of the documented sites, except for Canaleja 2 during the Early Neolithic. Here, the paleoenvironmental record does not present significant differences from what have been described for contemporaneous settlements like Los Barruecos o Cerro de la Horca (Cerrillo-Cuenca, et. al., 2010), that is, the documentation of incipient agriculture and livestock breeding in a not very transformed natural environment. Unfortunately, due either to stratigraphic disturbances or lack of information in samples, no paleoenviromental information could be recovered from sites like Tío Republicano, Cueva del Valle or La Charneca, which might have been useful for providing a more detailed picture on the subsistence practices developed around the sites. Even the analysis of stable isotopes could not have been performed in the charred bones of Tío Republicano, which could have offered an interesting material for comparison with the older data from Canaleja 2.

At the margins of the Tagus, where Canaleja Gorge is located, caves and megalithic sites occupy a similar extent, although a detailed inspection reveals that the geological background is a determinant factor for the location of one or another kind of sepulchre. In this area, caves are developed in small limestones systems, around which megalithic sepulchres can be found in granite and schist areas, a few kilometres away. In the rest of the areas, the relationship between caves and megaliths is also uneven, although the difference between the distributions of both phenomena is particularly close in the areas where different kinds of geological background can be found. For example, the Cueva del Valle site is surrounded by granite boulders, which allowed for the 85 

 

(Cerrillo-Cuenca et. al., in press, Enríquez, 1990), which is also compatible with the kind of sites documented alongside the Tagus drainage (González, 2012). In other words, it could be feasible to connect the low visibility in landscape of burial caves and the unstructured burials with communities organised in small habitats with unapparent traces of hierarchization. In fact, ossuaries have been interpreted either in terms of social equality, or, as Díaz del Río has pointed out (2006), as a mechanism for dealing with the first traces of inequality. In this sense Thomas (2000: 657) has argued that points out, ossuaries may have not been passive and closed, but rather a way for continuous reassuring of the identity of the living, interpretation that is also shared by other scholars (Tomé & Oosterbeek, 2011). From the more empirical point of view of anthropological analysis, the possibility of dealing with open ossuaries has also been proposed in Copper Age Iberia (Weiss-Krejci, 2005).

The assessment of social behaviours seems to be an alternative way for understanding the insertion of burial caves within a landscape. One main barrier for the interpretation could be the lack of well-preserved stratigraphic context and anthropological analysis for determining the nature and magnitude of the interred populations. During the Early Neolithic, possibly following the previous tradition, inhumation seems to be related to single individuals, as it happens in Nossa Senhora das Lapas (Oosterbeek, 1993), or in other openair contexts of Early Neolithic Iberia. The documented practice at Canaleja 2 finds similarities in other contexts of Iberia, where aisled human remains were recovered from apparent domestic contexts (Bernabeu, et. al., 2001). In this case, this small shelter might have acted as a place for the natural defleshing of corpses. The collective burials seem to have emerged at the same time as collective megalithic burials, by the middle of 5th millennium cal BC, which is clearly stated at the caves from the centre and south of Portugal. Ossuaries seem to have appeared in Portuguese caves during the middle Neolithic era, as the absolute date from Canaleja 1 can confirm for Spanish drainage of Tagus. However, it is difficult to establish at which point of the sequence cremation arises in the funerary ritual. In her complete study about cremation in Iberian Prehistory, Weiss-Krejci (2005) accounted for cremation as early as in the Neolithic period in Cadaval cave (Oosterbeek, 1995). However, with the data that we possess at the present time, we can only recognise this practice in our study area during the Copper Age. The data is scarce for proposing a reconstruction of the rituals linked to cremation. However, in the cases of Tio Republicano and Cueva del Valle, where all the bones seem to have been cremated, a “ritual of incorporation” of individuals (Williams, 2012) can be considered, which is compatible with the archaeological evidence from others sites of the Iberian Peninsula (Weiss-Krejci, 2005).

However, it is also probable that the meaning and the function of the ossuaries deposited at caves might have changed along time, and consequently, the ideology and the political aims that ruled their formation. Social transformations in Southwest Iberia have been adverted through this span of time with different forms and intensities, although they seem not to have affected the repetition of this ritual through time. This fact can possibly highlight the ability of this funerary behaviour to mask social inequalities throughout Late Prehistory. 4. Final remarks In this paper we have presented current appraisal of evidence for the burial caves of the Extremadura region. The commented data suggests that burial caves (and rockshelters) are a prominent feature in the analysis of funerary prehistoric behaviours, at least in a sense that might have not been sufficiently valued in the analysis of funerary evidences. Mesolithic and Early Neolithic burials are a true novelty in the region, which collaborates in connecting historical processes from neighbouring areas of Iberia. In addition, with the available evidences we can assure that burial caves played a complementary role to megalithic monuments, at least from the ending of 5th millennium cal BC, as the date from Canaleja 1 proves. The analysis of burial caves and rock-shelters illustrates a lack of intensive research in the Extremadura region; that is being overcome with the publication of new anthropological data (Cerrillo-Cuenca & González, 2007, Cerrillo-Cuenca & González, 2011; Muñoz & Canals, 2008; Barca, 2007). The general adverted pattern is that these caves were used for interring small groups of individuals, probably not more than 15 in each ossuary.

Copper Age ossuaries can be seen as a whole entity where the personhood is dissolved (Lucas, 1996) and where the ossuary emerges as a new social product composed of the remains of different individuals (Cerrillo-Cuenca & González, 2011; Tomé & Oosterbeek, 2011). The apparently unordered disposition of grave goods in the interior of the funerary space can be perceived as a collective offering, which can perhaps reinforce the interpretation of the collectivization of the ritual. Seen from this perspective, ossuaries like Tío Republicano, whose minimal number of individuals is 11, are compatible with a social context where the individuals might be related to each other through kinship. The indifferent treatment of bones of adults, children and infants can perhaps support this idea. Moreover, this data is coherent with an environment composed of small settlements, which might have played an independent role from the socio-political dynamics studied in the Central areas of Guadiana River during the 3rd millennium cal BC (Hurtado, 2005). Thus, burial caves from the Guadiana River basin seem to be constant with a series of small settlements detected in both areas

Along the present text, we have presented the general lines of our interpretation, grounded on the few known contexts. A general observed trend is that the ossuaries deposited at burial caves grew more complex along the sequence, which included rituals as cremation during Copper Age. Kinship relationship might justify the kind of ritual observed and even the number of interred 86 

 

individuals, which is coherent with the information that the pattern of settlements suggests. Nonetheless, this kind of behaviour might mask more complex social mechanisms, as previously stated. The research in similar

archaeological contexts would provide additional basis for understanding what has now been discovered to be an essential type of burial for prehistoric populations in the region.

References ALGABA SUÁREZ, M.; COLLADO GIRALDO, H.; FERNÁNDEZ VALDÉS, J. M. (2000) – Cavidades en Extremadura (España). Patrimonio natural y arqueológico. Oxford: BAR Publishing, BAR International Series 826. ÁLVAREZ, A. (1984) – Análisis de los restos óseos hallados en la cueva de Maltravieso (Cáceres). Revista de Estudios Extremeños. Badajoz: Centro de Estudios Extremeños, Tomo XL, nº 1 p. 171-176. ARAÚJO, A. C.; CAUWE, N.; SANTOS, A. I. (1995) – A necrópole neolítica (estudos das colecções das antigas escavações). In ARAÚJO, A. C.; LEJEUNE, M. – Gruta do Escoural: Necrópole Neolítica e Arte Rupestres Paleolítica. Lisboa: IPA, Trabalhos de Arqueologia, nº 8, p. 57-109. ARIAS, P.; CERRILLO-CUENCA, E.; ÁLVAREZ, E.; GÓMEZ-PELLÓN, E.; GONZÁLEZ, A. (2009a) – A view from the edges: the Mesolithic settlement of the interior areas of the Iberian Peninsula reconsidered. In MCCARTAN, S. B., SCHULTING, R., WARREN, G. & WOODMAN, P. C. (Eds.) – Mesolithic Horizons. Oxford: Oxbow Books, volume I, p. 303-311. ARIAS, P.; ARMENDARIZ, A.; BALBÍN, R.; FANO, M.; FERNÁNDEZ-TRESGUERRES, J.A.; GONZÁLEZ-MORALES, M.; IRIARTE, M.J.; ONTAÑÓN, R.; ALCOLEA, J.; ÁLVAREZFERNÁNDEZ, E.; ETXEBERRÍA, F.; GARRALDA, M. D.; JACKES, M.; ARRIZABALAGA, A. (2009b) – Burials in the cave New evidence on mortuary practices during the Mesolithic of Cantabrian Spain. In MCCARTAN, S. B.; SCHULTING, R.; WARREN, G.; WOODMAN, P. C. (Eds) – Mesolithic Horizons. Oxford: Oxbow Books, volume I, p. 650-656. BARCA, J. (2007) – Los restos óseos del yacimiento de Canaleja II, una pequeña contribución al estudio de las sociedades del Neolítico Antiguo. In CERRILLO CUENCA, E.; GONZÁLEZ CORDERO, A. – Cuevas para la eternidad: sepulcros prehistóricos de la provincia de Cáceres. Mérida: Asamblea de Extremadura, Ataecina, 3, p. 97-103. BARNATT, J.; EDMONDS, M. (2002) – Places apart? Caves and monuments in Neolithic and Earlier Bronze Age Britain. Cambridge Archaeological Journal. Cambridge: McDonald Institute for Archaeological Research, volume 12, issue 1, p. 113-29. BERNABEU, J.; MOLINA, L.; GARCÍA, O. (2001) – El mundo funerario en el horizonte cardial valenciano. Un registro oculto. Sagvntvm-PLAV. València: Departament de Prehistòria i Arqueologia de la Universitat de València, nº 33, p. 27-35. BEYNEIX, A. (2012) – Le monde des morts au

Néolithique en Aquitaine : essai de synthèse. L’Anthropologie. Paris: Elsevier, t. 116, nº 2, p. 222233. BLASCO, A.; EDO, M; VILLALBA, M.; SAÑA, M. (2005) – Primeros datos sobre la utilización sepulcral de la Cueva de Can Sadurní (Bergues, Baix Llobregat) en el Neolítico Cardial. Actas del III Congreso del neolítico de la Península Ibérica. Santander: Servicio de Publicaciones de la Universidad de Cantabria, p. 625-634. BLASCO RODRÍGUEZ, F.; ORTIZ ALESÓN, M. (1991) – Trabajos arqueológicos en “Huerta Montero”. Almendralejo, Badajoz. Extremadura Arqueológica II. Mérida-Cáceres: Junta de Extremadura, I Jornadas de Prehistoria y Arqueología en Extremadura (1986-1990), p. 129138. BOAVENTURA, R. (2011) – Chronology of Megalithism in South-Central Portugal. In GARCÍA, L.; SCARRE, C.; WHEATLEY, D. (Eds.) – Exploring Time and Matter in Prehistoric Monuments: Absolute Chronology and Rare Rocks in European Megaliths. Sevilla: Junta de Andalucía. Menga: Revista de Prehistoria de Andalucía, nº 1, p. 159-192. BOSCH-GIMPERA (1915-1920) – La cova del Boquique a Plasència. La cerámica. Anuari de l’Institut d’Estudis Catalans. Barcelona: Institut d’Estudis Catalans, MCMXV-XX, p. 514516. BUENO, P.; BALBÍN, R.; BARROSO, R. (2005) – El dolmen de Azután (Toledo). Toledo: Diputación de Toledo. BUENO, P.; BALBÍN, R.; BARROSO, R.; CERRILLOCUENCA, E.; GONZÁLEZ, A.; PRADA, A. (2011) – Megaliths and Stelae in the Inner Basin of Tagus River: Santiago de Alcántara and Cañamero (Cáceres, Spain). In BUENO, P.; CERRILLOCUENCA, E.; GONZÁLEZ CORDERO, A. (eds.) – From the Origins: the Prehistory of Inner Tagus Region. Oxford: BAR Publishing, BAR International Series 2219, p. 143-158. CALADO, M. (2000) – Neolitização e megalitismo no Alentejo central: una leitura espacial. Neolitização e megalitismo da Península Ibérica. Porto: ADECAP, Actas do 3º Congresso de Arqueología Peninsular, volume III, p. 35-45. CALLEJO, C. (1958) – La cueva prehistórica de Maltravieso junto a Cáceres. Cáceres: Publicaciones de la Biblioteca Pública de la Ciudad. CARVALHO, A.F. (2007) – Algar do Bom Santo: a research project on the Neolithic populations of Portuguese Estremadura (6th-4th millennia BC). Promontoria. Faro: Universidade do Algarve, nº 5, p. 185–198. 87 

CERRILLO-CUENCA, E. (2008) – Hábitats y ámbitos funerarios de la Prehistoria Reciente de Cáceres: El Conejar, Maltravieso y otros hallazgos aislados. In SANABRIA, P. J. – Investigaciones e intervenciones recientes en la ciudad de Cáceres y su entorno. Cáceres: Museo de Cáceres, Memorias, actas de las jornadas de arqueología del Museo de Cáceres, nº 7, p. 57-81. CERRILLO-CUENCA, E. (2009) – Láminas de sílex en el actual territorio de Extremadura (IV-III milenio CAL BC): problemas de partida y posibilidades de estudio. In GIBAJA, J. F.; TERRADAS, X. (coord.) – Les grans fulles de sílex a Europa al final de la prehistòria: actes. Barcelona: Museu d’Arqueologia de Catalunya, p. 55-62. CERRILLO-CUENCA, E. (2011) – Planteamientos y nuevos datos para la interpretación de los paisajes prehistóricos del sector extremeño del Tajo: el área de Alconétar. Zephyrus. Salamanca: Universidad de Salamanca, volume 68, nº 2, p. 139-161. CERRILLO-CUENCA, E.; CAZORLA, R.; CELESTINO, S.; SALAS, E. (in press) – Comunidades prehistóricas y sepulcros colectivos en cueva en la Cuenca del Guadiana. A propósito de de los materiales de la cueva del Valle (Zalamea de la Serena, Badajoz). VI Jornadas de Arqueología del Suroeste Peninsular. CERRILLO-CUENCA, E.; GONZÁLEZ, A. (2007) – Cuevas para la eternidad: sepulcros prehistóricos de la provincia de Cáceres. Mérida: Asamblea de Extremadura, Ataecina, 3. CERRILLO-CUENCA, E.; GONZÁLEZ, A. (2011) – Burial prehistoric caves in the interior basin of River Tagus: the complex at Canaleja Gorge (Romangordo, Cáceres, Spain). In BUENO, P.; CERRILLOCUENCA, E.; GONZÁLEZ CORDERO, A. (eds.) – From the Origins: the Prehistory of Inner Tagus Region. Oxford: BAR Publishing, BAR International Series 2219, p. 23-42. CERRILLO-CUENCA, E.; GONZÁLEZ, A.; HERAS, F. J. (2009) – Cuevas funerarias en el Tajo Interior: a propósito de Maltravieso. In SANABRIA, P. J. – Actas del Congreso El Mensaje de Maltravieso 50 años después (1956-2006). Cáceres: Museo de Cáceres, Memorias, nº, 8, p. 209-222. CERRILLO-CUENCA, E.; GONZÁLEZ, A.; LÓPEZ, J. A.; LÓPEZ, L. (2010) – La primera mitad del Holoceno en el territorio de Extremadura: datos arqueológicos y paleoambientales. In GIBAJA, J. F.; CARVALHO, A. F. (eds.) – Os ultimos caçadores recolectores e as primeiras comunidas produtoras do sul da Península Ibérica e do norte de Marrocos. Faro: Universidade do Algarve, Promontoria Monográfica, nº 15, p. 81-88. CHAMBERLAIN, A. (1996) – More Dating Evidence for Human Remains in British Caves. Antiquity. Durham: Durham University, volumen 70, nº 270, p. 950–53. COLLADO, H.; FERNÁNDEZ, M.; POZUELO, D.; GIRÓN, M. (1997) – Pinturas rupestres esquemáticas en la transición del IV al III milenio a.C. El abrigo de la Charneca Chica (Oliva de Mérida, Badajoz).

Trabajos de Prehistoria. Madrid: CSIC, volume 54, nº 2, p. 143-149. COLLADO, H.; GARCÍA, J. J. (2010) – 10000 años de arte rupestre. El ciclo preesquemático de la Península Ibérica y su reflejo en Extremadura. Global Rock ART. Anais do Congreso Internacional de Arte Rupestre IFRAO (Parque Nacional Serra da Capivara, Piauí, Brasil 29 de Junho a 3 de Julho de 2009). São Raimundo Nonato: FUMDHAM Fundação Museu do Homem Americano, Fumdhamentos, volume 4, nº 9, p. 1168-1192. CRUZ, A. (1997) – Vale do Nabão. Do Neolítico à Idade do Bronze. Tomar: Centro Europeu de Investigação da Pré-História do Alto Ribatejo, Arkeos, perspectivas em diálogo, volume 3, 361 p. DIAZ-ZORITA, M.; COSTA, M. E.; GARCÍA, L. (2012) – Funerary practices and demography from the Mesolithic to Copper Ages in Southern Spain. In GIBAJA, J.F.; CARVALHO, A.; CHAMBON, P. (Eds.) – Funerary Practices in the Iberian Peninsula from the Mesolithic to the Chalcolithic. Oxford: BAR Publishing, BAR International Series 2417, p. 51-65. DÍAZ DEL RÍO, P. (2006) – An appraisal of social inequalities in Central Iberia (c. 5300-1600 cal BC). In DÍAZ DEL RÍO, P.; GARCÍA SANJUÁN, L. (Eds) – Social Inequality in Iberian Late Prehistory. Oxford: BAR Publishing, British Archaeological Reports International Series 1525, p, 69-79. DOWD, M. (2008) – The Use of Caves for Funerary and Ritual Practices in Neolithic Ireland. Antiquity. Durham: Durham University, volumen 82, nº 316, p. 305-317. ENRÍQUEZ NAVASCUÉS, J. J. (1986) – Excavación de urgencia en la cueva de la Charneca (Oliva de Mérida, Badajoz). Noticiario Arqueológico Hispánico. Madrid: Ministerio de Cultura, nº 28, p. 7-24. ENRÍQUEZ NAVASCUÉS, J. J. (1990) – El Calcolítico o Edad delCobre de la cuenca extremeña del Guadiana: los poblados. Badajoz: Publicaciones del Museo Arqueológico Provincial de Badajoz. ENRÍQUEZ NAVASCUÉS, J.J.; RODRIGUEZ DÍAZ, A. (1990) – Algunos ídolos en barro cocido y hueso de la Baja Extremadura. Zephyrus. Salamanca: Universidad de Salamanca, volume 43, p. 101-107. ESPARZA ARROYO, A. (1990) – Sobre el ritual funerario de Cogotas I. Boletín del Seminario de Arte y Arqueología. Valladolid: Universidad Valladolid, LVI, p. 106-143. ETXEBERRÍA, F.; DELIBES, G. (2002) – Interpretación del fuego en los sepulcros megalíticos. In Rojo-Guerra, M.; Kunst, M. (Eds) – Sobre el significado del Fuego en los rituales funerarios del Neolítico. Valladolid: Universidad Valladolid, Studia Archaeologica, 91, p. 59-64. GARCÍA, J.J.; COLLADO, H.; FERNÁNDEZ, M.; GIRÓN, M.; GARCÍA, M.I.; MESA, M.J. (2011) – La Cueva Chiquita o de Álvarez (Cañamero, Cáceres): Recientes intervenciones y revision de sus manifestaciones rupestres. Espacio, Tiempo y Forma. Madrid: UNED, Prehistoria y Arqueología, 4, Serie I, Nueva época, p. 81-109. 88 

GARCÍA, P. (1915-1920) – La cova del Boquique a Plasència. L’excavació i les troballes. Anuari de l’Institut d’Estudis Catalans. Barcelona: Institut d’Estudis Catalans, MCMXV-XX, p. 514-516. GARCÍA, P.; SALAZAR-GARCÍA, D.; PÉREZ, A.; PARDO, S.; CASANOVA, V. (2011) – El Neolítico antiguo cardial y la Cova de la Sarsa (Bocairent, València). Nuevas perspectivas a partir de su registro funerário. Munibe (Antropologia-Arkeologia). San Sebastian – Donostia: Sociedad de Ciencias Aranzadi, nº 62, p. 175-195. GONÇALVES, V. S. (1978) – A Neolitização e o Megalitismo da região de Alcobaça. Lisboa: Secretaria da Cultura. GONZÁLEZ, A. (2012) – La Edad del Cobre en la Alta Extremadura. Asentamientos y organización del territorio. Cáceres: University of Extremadura. Unpublished doctoral disertation. GONZÁLEZ, A.; RASCÓN, J. (Unpublished) – Estudio antropológico de los restos óseos humanos recuperados de la cueva de Tío Republicano (Romangordo, Cáceres). HIERRO, J. A. (2011) – La utilización sepulcral de las cuevas en época visigoda: los casos de Las Penas, La Garma y el Portillo del Arenal (Cantabria). Munibe (Antropologia-Arkeologia). San Sebastian – Donostia: Sociedad de Ciencias Aranzadi, nº 62, p. 351-402. HURTADO, V. (2005) – El Campaniforme en Extremadura. Valoración del proceso de cambio socioeconómico en las cuencas medias del Tajo y Guadiana. In ROJO GUERRA, M.A.; GARRIDO PENA, R.; GARCIA MARTINEZ DE LAGRAN I. (Coord.) – El Campaniforme en la Península Ibérica y su contexto europeo/Bell beakers in the Iberian Peninsula and their european context. Valladolid: Universidad Valladolid, p. 321–335. HURTADO, V.; MONDÉJAR, P.; PECERO, J. C. (2000) – Excavaciones en la tumba 3 de La Pijotilla. In JIMÉNEZ ÁVILA, J.; ENRÍQUEZ NAVASCUÉS, J. J. (Eds.) – El Megalitismo en Extremadura. Homenaje a Elías Diéguez Luengo. Mérida: Junta de Extremadura, Extremadura Arqueológica, VIII, p. 249-266. JORDÁ, J. F.; AURA, J. E. (2008) – 70 fechas para una cueva: revisión crítica de 70 dataciones C14 del Pleistoceno Superior y Holoceno de la Cueva de Nerja (Málaga, Andalucía, España). Espacio, Tiempo y Forma. Madrid: UNED, Prehistoria y Arqueología, 1, Homenaje al Profesor Eduardo Ripoll Perelló, Serie I, p. 239-256. JORDA, J. F.; MESTRES, J. S. (1999) – El enterramiento calcolítico precampaniforme de Jarama II: una nueva fecha radiocarbónica para la Prehistoria Reciente de Guadalajara y su integración en la cronología de la región. Zephyrus. Salamanca: Universidad de Salamanca, volume 52, p. 175-190. LEISNER, G.; LEISNER, V. (1951) – Antas do concelho de Reguengos de Monsaraz (Materiais para o estudo da cultura megalítica em Portugal). Lisboa: Instituto para a Alta Cultura.

LUBELL, D.; JACKES, M.; SCHWARCZ, H.; KNYF, M.; MEIKLEJOHN, C. (1994) – The MesolithicNeolithic transition in Portugal: isotopic and dental evidence of diet. Journal of Archaeological Science. Canada: Academic Press, volume 21, issue 2, p. 201-216. LUCAS, G.M. (1996) – Of death and debt: a history of the body in Neolithic and Early Bronze Age Yorkshire. Journal of European Archaeology. Praga: European Association of Archaeologists, nº 4, p. 99118. MARTÍNEZ, M. I. (1984) – El comienzo de la metalurgia en la provincia de Madrid. La cueva y el cerro de Juan Barbero (Tielmes). Trabajos de Prehistoria. Madrid: Consejo Superior de Investigaciones Científicas, volume 41, nº 1, p. 17128. MUÑOZ, L.; CANALS, A. (2008) – Nuevos restos humanos hallados en la cueva de Maltravieso. In SANABRIA, P. J. – Actas del Congreso El Mensaje de Maltravieso 50 años después (1956-2006). Cáceres: Museo de Cáceres, Memorias, nº, 8, p. 205208. OOSTERBEEK, L. (1993) – Nossa Senhora das Lapas: excavation of prehistoric cave burials in central Portugal. Papers from the Institute of Archaeology. Oxford: Institute of Archaeology, nº 4, p. 42-64. OOSTERBEEK, L. (1997) – Back home! Neolithic life and the rituals of death in the Portuguese Ribatejo. In BONSALL, C.; TOLAN-SMITH, C. (ed.) – The human use of caves. Oxford: BAR Publishing, British Archaeological Reports International Series 66, p. 70-80. PAN, I. (1917) – Exploración de la cueva prehistórica del Conejar (Cáceres). Boletín de la Real Sociedad Española de Historia Natural. Madrid: Museo Nacional de Ciencias Naturales, XVII, p. 185-190. PELLICER, M.; ACOSTA, P. (1982) – El Neolítico Antiguo en Andalucía Occidental. Le Neolithique Ancien Mediterranéen. Archaeologie en Languedoc, 0. Especial, p. 49-60. PIÑÓN, F. Y BUENO, P. (1985) – Los grabados del sepulcro megalítico de Magacela, (Badajoz). Tres estudios sobre el Calcolítico extremeño. Cáceres: Universidad de Extremadura, Departamento de Prehistoria y Arqueología, Series de Arqueología Extremeña, 1, p. 65-81. PRADA, A.; CERRILLO-CUENCA, E. (1996-2003) – Megalitismo y poblamiento neolítico en el suroeste de Badajoz: una lectura complementaria. Norba Historia. Cáceres: Universidad Extremadura, p, 4774. RUIZ-GÁLVEZ, M. (2000) – El conjunto dolménico de la Dehesa Boyal de Montehermoso. In JIMÉNEZ, J.; ENRÍQUEZ NAVASCUÉS, J. J. (Eds.) – Homenaje a Elías Diéguez Luengo. Mérida: Extremadura Arqueológica, VIII, p. 187-207. SCARRE, C. (2009) – Stony ground: outcrops, rocks and quarries in the creation of megalithic monuments. In SCARRE, C. (Ed.) – Megalithic Quarriying. Sourcing, extracting and manipulating the stones.

89 

Oxford: BAR Publishing, British Archaeological Reports International Series 1923, p. 3-20. THOMAS, J. (2000) – Death, identity and the body in Neolithic Britain. Journal of the Royal Anthropological Institute. London: Royal Anthropological Institute, 6, p. 603-17. TOMÉ, T. (2010) – Até que a Morte nos Reúna: Transiçao para o Agro-Pastoralismo na Bacia do Tejo e Sudoeste Peninsular. Vila Real: University of Trás-Os-Montes e Alto Douro. Unpublished doctoral dissertation. TOMÉ T.; OOSTERBEEK, L. (2011) – One region, two systems? A Paleobiological Reading of Cultural Continuity over the Agro-Pastoralist Transition in the North of Ribatejo. In BUENO, P.; CERRILLOCUENCA, E.; GONZÁLEZ CORDERO, A. (eds.) – From the Origins: the Prehistory of Inner Tagus Oxford: BAR Publishing, BAR Region. International Series 2219, p. 43-54. UBELAKER, D. (2009) – The forensic evaluation of burned skeletal remains: A synthesis. Forensic Science International. Oxford: Elsevier, volumen 183, issue 1-3, p. 1-5.

VIDAL, J. M.; PRADA M. E.; FUERTES, M. N.; FERNÁNDEZ, C. (2010) – Los hombres mesolíticos de la Braña/Arientero (Valdelugueros, León): El hallazgo, situación aspectos arqueo-antropológicos, cronología y contexto cultural. In VIDAL ENCINAS, J.; PRADA MARCOS, M. E. (Coord) – Los hombres mesolíticos de la cueva de la BrañaArintero (Valdelugueros, León). León: Junta de Castilla y León/Consejería de Cultura y Turismo/Museo de León, Estudios y catálogos, 18, p. 18-61. WILLIAMS, D. (2012) – Towards an Archaeology of Cremation. In SCHMIDT, C. W.; SYMES, S. A. (Eds.) – The Analysis of Burned Human Remains. New York: Elsevier Press, p. 239-270. WEISS-KREJCI, E. (2005) – Formation processes of deposits with burned human remains in Neolithic and Chalcolithic Portugal. Journal of Iberian Archaeology. Porto: Universidade do Porto, nº 7, p. 37-73. ZILHÃO, J. (1992) – Gruta do Caldeirão – o Neolítico Antigo. Lisboa: IPPAR, Trabalhos de Arqueologia 6.

90 

Between dead and alive - the recent prehistory of the municipality of Pampilhosa da Serra (Central Portugal) Carlos Batata, Filomena Gaspar Situados uns e outros, em zona muito montanhosa, é nas cabeceiras do Rio Unhais que a arte mais se manifesta, em estreita relação com alguns destes monumentos funerários. Foram efectuados trabalhos arqueológicos em algumas destas mamoas, nos anos de 2009 a 2011, alguns desenhos de motivos artísticos e ensaiadas técnicas de limpeza das rochas, por métodos naturais.

Abstract The realization of the survey for edition of the Carta Arqueológica do Concelho de Pampilhosa da Serra (edited in April 2009) has revealed, to the moment, 141 “mamoas”, tumuli or little tumulus, associated with about 151 panels of rock art. The most part of this art could be inserted in the prehistoric age, but there is also some manifestations of recent periods like the Bronze Age. Both implanted in very mountainous places, it is in the water springs of the Unhais River that is found the most part of the rock art, in straight relation with some of these funerary monuments. Archaeological work was done on some of these tumuli, in the years 2009 to 2011, some drawings of artistic motifs and was tested cleaning techniques of the rocks, with natural methods.

Introduction Upon completion of the first archaeological survey of the municipality of Pampilhosa da Serra (Batata & Gaspar, 1994), hasn’t been found any megalithic monuments, in despite of the registering of some place names as Ucheiros or Vale das Antas. At the time, the characteristics of these small tumuli were still less known in archaeological community. It was also verified in the archaeological bibliography available, that such structures have been identified and explored, especially those who were near dolmens. However, they were not as important as they are today, because the research projects of the time were more directed to the study of dolmens. These were clearly visible in the landscape of Alentejo, Beiras, and Northern Portugal, while these mounds had little volumetric expression, often passing unnoticed or regarded as clusters of rocks from the cleaning of farm lands. The characteristics of these small tumuli (lots of small stones forming a low height mound) have not caused in local populations the creation of legends or stories. For them, these mounds were only stone mounds (figure 8.1), that they knew but gave them no strange character, like the work of ancient people.

Resumo A realização da prospecção arqueológica para a edição da Carta Arqueológica do Concelho de Pampilhosa da Serra (editada em Abril de 2009) e o desenvolvimento do projecto de investigação Arte rupestre e monumentos funerários da Pré-história recente do concelho de Pampilhosa da Serra, revelaram, até ao momento, a existência de 143 mamoas, tumuli ou montículos, associados a cerca de 150 painéis com arte rupestre. Nos concelhos de Fundão, Sertã, Vila de Rei, Sardoal e Mação localizámos ainda algumas mamoas com as mesmas características. Uma boa parte desta arte, e algumas mamoas, podem inserir-se no período NeoCalcolítico, embora também haja muitas manifestações de períodos mais recentes como a Idade do Bronze.

Figure 8.1: Mamoa VI de Vilares, Pampilhosa da Serra. 91 

 

possible remains of a defensive wall (Dias, 1985). In field exploration, we have found a settlement with abundant pottery, part of it from the Final Bronze Age and a Bronze axe fragment of amphibolite (Batata & Gaspar, 1994: 36-40); Cabeço Redondo was then classified as a Bronze Age settlement, and today believed to be of an older occupation. In the Serra do Machialinho, were found by villagers and acquired for the municipality, a set of spiral gold rings, weighing 37.2 gr. (Batata & Gaspar, 1994: 48-52), which context wasn’t possible to ascertain. Still in the same mountain (Cova da Iria), we found a retouched flint blade fragment, in an area that was very churned up, had we, at the time, doubted about their archaeological context (Idem, Ibidem: 52).

On the other hand, the first news about rock art were obtained after the completion of the archaeological survey in the municipality of Pampilhosa da Serra. During the field survey, the still alive Filomena Gaspar’s maternal grandfather had taken us to see the “footprint of Saint Mary” which, unfortunately, was a natural cavity, with the approximate format of a foot, on top of some rocks. He also referred to us the existence of engravings in rocks, in Serra das Fontes (in the parish of Dornelas do Zêzere) that we never came to visit in his company. Another familiar of her, spoke to us about the engravings of Vale do Gato (near the village of Malhada do Rei) that he had observed when he tended cattle in his times of youth. We visited and photographed them and did our withdrawal with plastic crystal sheets in 1996, which was p ut in second plan until it was integrated on the Carta Arqueológica do Concelho de Pampilhosa da Serra, published in 2009. In 2000, in an article about the rock art of Pinhal Interior’s area, we made a first characterization of the engravings in Vale do Gato (Batata & Gaspar, 2000: 575-581).

With regard to rock art, it is prominently found at the west of the municipality of Pampilhosa da Serra, as is the case of Pedra Letreira (≈ engraved stone) (Cabeçadas) and Mestras I, II and III (Pedra Riscada) (Nunes, et. al., 1974: 6-25) and at the south (Serra do Cabeço Raínho, Oleiros and Sertã) (Batata, 1998, p. 18-21). To the North, in the area of Piódão were also studied some rock art panels, where some patterns resemble already studied patterns (Santos, et. al., 2011: 161-176).

Geomorphological Environment The municipality of Pampilhosa da Serra is a municipality with deep valleys and high mountains. It can be divided into two geographical areas which, although they belong to the same geological formation, have some morphological differences. Thus, the High County, which covers mostly the parishes of Cabril, Vidual, Unhais-oVelho and Fajão, features mainly quartzitic rock outcrops, inserted through the mountains, with generation at the base of the Silurian-Ordovician (Thadeu, 1951: 1170). The so-called Low County covers the rest of the county and does not show noteworthy reliefs. Geologically speaking consists of clayish shale that, contrary to what happens on the High County, that has rounded ridges and mountains of the pterozoic system. Between these mountainous systems there is a wide hydrographic network, which seeps through the rivers Ceira, Unhais and Zêzere. These, on the other, run between tight valleys, in a V shape, sketching, in its way, huge meanders.

To the East of the County, was printed, in 2002, a summary of archaeological sites and materials, if not helped us define the prehistory of the county in 1994, are important to contextualize the set of burial mounds and rock art identified in 2009. So, there is a settlement of the Final Bronze, with posterior occupation, situated in Cabeço da Argemela (Lavacolhos, Fundão) and two tumuli of the same type as in Pampilhosa da Serra, in Quinta da Caneca, in Cova da Beira, and a bronze deposit in Paúl (Covilhã) (Vilaça, et. al., 2000: 200-210, 214 and 217). Regional Context It’s begining to be plenty the sets of tumulus, mounds or tumuli (according to the designation assigned to them by investigators who dedicated themselves to the study of this type of funerary monuments) that are scattered or grouped into necropolis, throughout the country, with special emphasis on its northern half.

Pre-existent Prehistorically data In the surrounding area, to the date of our initial work, we’ve collected the information available in the bibliography that showed the existence of a prehistoric Idol, appeared in a low area, next to the river Ceira and the village of Relvas de Teixeira (Arganil), located on the right river bank and in front of Fajão (village situated on the left bank), on which we would come to find a prehistoric settlement that at the time, we dated as belonging to the Bronze age, based only on a handful of a quite rude amorphous manual ceramic fragments (Batata & Gaspar, 1994: 21) and the findings of Relvas (a prehistoric anthropomorphic Idol, in quartzite, with 18 cm in diameter and 23 high and an bronze axe) (Nunes, 1956: 503-507). In a prominent mound suggestively called Cabeço Murado (= walled mound) (Trinhão), it was found a fishing net plummet, associated with the

One of the oldest, or in other words, one of which has been excavated for a longer time, it is the 2nd Tumulus of Chã de Carvalhal (Serra da Aboboreira, Baião) (Cruz, 1990: 151-155), situated in the vicinity of a dolmen. Deserving the reference, there is a group of tumuli with the same characteristics, the set of 40 monuments of Maciço da Gralheira, situated between the Serra de Montemuro and the Serra do Caramulo (Silva, 1997: 605620). Closer in time and space, other researchers (Cruz, et. al., 2000: 125-150) have identified and excavated some burial mounds of the same type as the ones in Pampilhosa da Serra. In the area of Vila Nova de Paiva were identified dozens of monuments. 92 

 

At Quinta da Caneca (Fundão), in Cova da Beira, were identified two burial mounds similar to the ones in Pampilhosa da Serra (Vilaça et. al., 2000, p. 200-2001, 214 and 217).

As to the state of conservation of monuments (figure 8.3), it turns out that the 141 tumuli are divided into very similar amounts between the well preserved monuments and the poorly preserved monuments.

In the municipality of Vila Pouca de Aguiar, the authors (Batata, et. al., 2008: 98-99) also identified two funerary structures that present similarities with these monuments. During the archeological survey carried out in 2008, in the municipality of Amarante, a signatory also identified two burial mounds (Batata, 2008: 60).

As for the altimetric positioning on the field (figure 8.4), about 40% of tumuli are situated in quotas between 650 and 800 m, which correspond roughly to the ridges of mountains.

In counties closest to Pampilhosa da Serra also has been identified such monuments. In the municipality of Sabugal were identified by these researchers (Caninas, et. al., 2009: 21-38) a dozen of small and large burial mounds, a few dozen in Oleiros and some in the serra da Lousã (Caninas, et. al., 2004: 3-23; Caninas, et. al., 2008: 89-184), with archaeological excavation of some of them (the Vale de Mós I, Feiteira and Selada do Cavalo). In Vila Velha do Rodão, in the context of review of the PDM were identified 13 tumuli, all grouped in the same location (Henriques, et. al., 2008: 79-88).

Figure 8.2: Number of tumuli according to the diameter.

Brief Characterization of the Burial Mounds The fact that the mountains had some fires and it was burned almost all the vegetation, helped in large part to see a significant amount of tombs that would otherwise be much more difficult to find. In a generic way, burial mounds appear across the county, with the exception of the South, next to the Zêzere River, which is a very precipitous area (between Lobatinhos, Signo Samo and Portela do Fojo). Looking at the chart (figure 8.2), it turns out that small tumuli, associated or not to larger ones, represent a significant percentage of the total of burial mounds identified on the field. These burial mounds are very visible on the ground, by featuring a stone carapace, mostly composed by medium and small blocks of milky quartz, very abundant in this area, due to the quartzitic mountains, impregnated with milky quartz veins. There are, however, mounds composed exclusively with fragments of shale, but these are a minority. There is also the case of a small mound, with 3 m in diameter is made only with dirt. The reason for this to happen is that its builders used the raw material that existed on the site. In rich milky quartz areas, the tumuli are almost exclusively composed with this material. In places where there are no veins of quartz, is used the local raw material. By that, there were burial mounds with stone carapace in shale, burial mounds with shell made with pebbles of quartzite (those that are closest to rivers or streams) and a number of tumuli made with fragments of angular quartzite. This last set, situated in Serra do Machialinho and the place where we found a fragment of a retouched flint blade (Cova da Iria) and pebbles from the Zêzere river, is very disformed by pine plantations.

Figure 8.3: Number of tumuli according to its state of conservation.

Figure 8.4: Number of tumuli in the altitude function.

93 

 

There is still a good percentage to lower quotas (35%), in the slope of the mountains or in small hills. As elsewhere, here too, a good percentage of burial mounds (about 35%) were deployed in platforms or slopes, meaning that the visibility intended to give to tumuli, was not only done preferably at the top of the hills, but also in places of passage or transit (figure 8.5). In virtue of the natural elements and also by anthropic reasons, the current height of the tumuli is not equal to that which would be when they were built. The stone carapaces and dirt forming the tumulus yield to natural elements such as water, ice and wind, causing its drawdown. On the other hand, the violation of the burial chamber, with the purpose of scavenging treasures or parts also contributed to its lower or raze, in many cases (figure 8.6).

Figure 8.5: Number of tumuli on the basis of topographic deployment.

Archaeological Work Since the publication of the Carta Arqueológica its authors intervened in three funerary monuments quite run-down, which had the risk of disappearing completely. These interventions, in addition to the preventive nature of the excavation, provided some scientific data as regards the typology of monuments and their chronology. Tumulus of Cabeço da Linteira Figure 8.6: Number of tumuli in function of its height.

The tumulus of Cabeço da Linteira is situated in the middle of a forest path, from it only remaining its lithic ring, since the passage of the leveller destroyed the stony carapace composed with milky quartz and shale, which are scattered along the way (figure 8.7).

Figure 8.7: Tumulus of Cabeço da Linteira, after excavation.

94 

 

Figure 8.8: Plant of what remains of the Tumulus of Cabeço da Linteira. What remained of the monument was the lithic ring of the funerary monument, forming a circle nearly perfect, with 6.80 m in east-west direction and 6.60 m in north-south direction, 50 cm in width, and thickness of 10 cm to the southwest and 20 cm to the to the northeast, with no signs of the burial chamber. On the north side, in a small extension, it was verified the existence of brown dirt, over the lithic ring, that would be the coverage of the monument, and some schist slabs, belonging to the stony carapace.

lithic materials accompanying the burial, then perhaps they have origin in a habitat that exist not much far away. 5th Tumulus of Vilares The 5th tumulus of Vilares is located on a very little hill, between two minor water lines. Is part of a set of six tumuli, grouped two at a time, being a larger and a smaller, forming a lined necropolis and with two pairs in small hills and one in a slope that lies between the other sets.

The monument is implanted on a hill, with a slight inclination to the Northeast, to 770 m in quota with a visual field of landscape, situated between the Zêzere river to the South and Unhais River to the North not far from it runs a pre-Roman route, but also used by Roman and Medieval people. This route is about 500 m south of the tumulus at the base of the burial mound of Vale Serrão, with a stony carapace which presents many similarities with this tumulus.

The excavation of this tomb structure revealed the existence of a small tumulus with 2 m in diameter, with a lithic irregular ring (figures 8.9 and 8.10), almost circular and made of large blocks of milky quartz. Part of the lithic ring, on the north side, was missing, with the stone removed by the owner to build a house at about 25 years ago.

Due to the high state of destruction of the monument, we cannot determine whether the ceramics and stone artifacts found within the lithic ring and in the little that was left of the U.E [2] (which was the land cover), were preexisting materials to the monument or if had been packed in the dirt transported to form the tumulus. If they were transported and have nothing to do with any ceramic and

It had a little decentered hole, and all the internal space was covered with a layer of shale, target of violation at an uncertain date. It would contain a pot or a bowl, from which we have found some amorphous fragments. On this layer of shale was placed a carapace constituted solely by blocks of milky quartz of medium and small size. Although we have not chronology for the funerary monument, at first glance seems to be a late monument, perhaps from the Bronze Age.

95 

 

Figure 8.9: Aspect of the 5th Tumulus of Vilares, after excavation.

Figure 8.10: Plant of what remains of the 5th Tumulus of Vilares.

96 

 

6th Tumulus of Vilares

dispersion dimension of the milky quartz and some shale that constituted the carapace of the tumulus. This cleaning permitted the delimitation of the surface stones, and to visualize other ones that were covered by the humous layer, for us to be able to register them.

The tumulus was detected during the survey carried out for the preparation of the Carta Arqueológica do Concelho de Pampilhosa da Serra (Batata & Gaspar, 2009: 108) and is part of a set of 143 tumuli detected in physical space of the county (data from August 2009). It is 18 meters southwest of the 5th Tumulus of Vilares who suffered an emergency intervention in the year 2009.

The general cleaning of the carapace revealed the existence of some reddish pottery fragments, found on the surface or in the EU [1]. It all appeared in squares in the lithic containment ring of the carapace. The cleaning also allowed the better seeing of the lithic ring, consisting of large blocks of milky quartz, forming a nearly perfect ring.

The tumulus was divided into four quadrants (figure 8.11) and within those quadrants was open squares with 1 x 1 m, north-oriented and numbered with a letter in longitude and in latitude an Arabic number. In order to obtain the profiles north-south and east-west, were dug to the bedrock, two ditches with 50 cm wide, signalized on the checkered grid of the mound.

It was done the drawing of all the stones of the tumulus, being clearly visible the location of the burial chamber through a small central depression and a greater concentration of shale fragments.

The tumulus was covered with shrub and young pine trees, born after the great fires in the area a few years ago (figure 8.1). The pine trees that were on the carapace of the tumulus were cut, with the owner’s authorization, in order to prevent future destruction by these large trees. The shrubby vegetation, consisting of genista, heather and gorse has been removed.

Then we proceeded to the removal of the smaller stones in the surface that were loose and scattered on the mound, as well as the full removal of all stones outside the lytic ring. Some of them resulted from the falling of stones in the covering area over time, other from stone extraction for construction, as has been confirmed by the owner: this tumulus, the 5th tumulus of Vilares and the 4 remaining burial mounds that are part of this necropolis, left several loads of stone for construction of the owner’s houses, about 25 years ago

After the cleaning and the making of the grid, we proceed to the removal of the humous layer consisting of leaves, mosses, roots and some dirt, to obtain the maximum

Figure 8.11: Final design of the 5th Tumulus of Vilares. 97 

 

Figure 8.12: Recovery of 6th Tumulus of Vilares. The excavation was processed from the outside to the inside, that is, we began to excavate the squares outside the lytic ring, to the rock. All the dirt was riddled. It was found that some large stones of lytic ring were missing, what should have happened at the time of extraction of stone for construction. It was found that, although the lytic ring form a near-perfect circumference (diameter with 4.10 m East-West and North-South with 4, 20 m), the layout of the blocks is relatively uneven. It was also verified that the lytic ring was settled directly on the rock and that in some places, were placed a few small milky quartz stones to lay the large blocks. In the southern part, due to the slight slope of the land, the amount of these small stones became bigger. In some of the exterior squares were found large chunks of coal, placed on the shale outcrop.

aside the chance of a burnt level prior to the construction of the tumulus. It only remained the hypothesis of coals as the result of forest fires after the construction of the funerary monument. Carbon-14 dating revealed that the coals dated from the 11th and 12th centuries. The excavation of the ditches allowed to know that the first layers of the carapace were formed by large blocks of milky mixed with blocks of shale and greywacke, with missing stones. We left to the end the excavation of the burial chamber, located at the intersection of the two ditches. When we made the initial drawing, it was found that there was a depression in the center of the burial mound, and that the amount of shale chips was more abundant in this central area. The excavation area delimited an irregular format violation crater and deep excavation revealed some large stones that had been pulled out by looters or for extraction of stones for the construction of houses. Some of the larger blocks suggest the existence of a stone circle that would be the burial chamber. The state of destruction does not allow being very conclusive. There was also the hole of the burial chamber, dug in the soil, in the shape of a crescent. It wasn’t found any fragment of pottery or lithic material. As there was also coal inside the chamber, the looting of this monument can only have occurred after the 12th century.

It was also found that, during the preparations for the construction of the tumulus, some vertical shale outcrops were deliberately flattened. We don’t know if within the tumulus there were also flattened outcrops, given that has not been done the dismantling of the entire structure. In fact, was only made the internal dismantling of a ditch of 50 cm wide by 3, 5 m long, on the north side and another in the West-East direction, with the same width and length of 3 m. This ditches were made to clarify how was formed the tumulus, having been dug down to the bedrock, and whether there were coals inside that could provide a secure chronology, since the coals collected outside the tumulus, didn’t meant that they could be dated at the time of construction of the tumulus .

Conservation of the 6th Tumulus of Vilares In order to prevent its destruction and constitute an educational item, we contacted the parish of Unhais-oVelho and municipality of Pampilhosa da Serra, in a way to preserve the monument. The first steps have been

After removing all the stones of the carapace in the ditches, it was found that there was no coal inside, putting 98 

 

In this excavation, as we said, we collect outside the tumulus some coals that do not appear in secure informations. We also collected coals inside the looted chamber that date to a post quem loting, but not the dating of the monument.

taken, through the placement of geotextile and a layer of sand over this, followed by the replacement of the quartz stones which had been taken from the carapace of the tumulus and spread around. We proceeded to fill the gaps of the lithic ring, through the placement of stones that were missing. The excavation left an area with about 2 m wide around the monument, which may take a layer of gravel to prevent soil erosion caused by visitors. Such work is not done, and so the Serra-Mãe association will request to the Town Hall and Parish council the placement of the gravel, with the accompaniment of an archaeologist of the association, as well as an information panel and signboards.

In terms of comparison, if compared to the construction technology and materials collected from the tumulus of Cabeço da Linteira (parish of Pampilhosa da Serra) that we insert in the Chalcolithic, these two tumuli, contemporary to each other, should date a period later, insertable in the Bronze Age. As for the lithic materials collected, the large amount of shale caps is recurring in the area, lying almost everywhere. Shown to people in the area, they were not of any contemporary use, so we must consider them old, failing, however, to associate a use with these roughly made discs, with a round and oval format.

The excavation of the tomb structure revealed the existence of a tumulus about 4 m in diameter, with an irregular lithic ring carapace, such as the case in the 5th tumulus of Vilares that is close to this one. The central hole is also very similar to that found in this small tumulus of 2 m in diameter, as well as the coverage of the burial chamber with schist slabs. As the previous one this one was also looted in an uncertain date, with the only difference of having revealed fragments of archaiclooking ceramic, which did not happen in this tumulus. The carapace would be equal in two tumuli.

Cleaning and Design of Rock Art In 2011, we started the registering of the drawings engraved in the rocks and tried out some lichens cleaning techniques, through natural process observed on site, before looking to chemicals. In fact, we found that the rocks that were covered with dirt had the surfaces clean of any lichens: those already exposed, sometimes had large amount of them that made difficult the reading and drawing, and does not allowed taking good photos. The technique used consisted to cover some engravings, in that the stone was blackened and with many lichens, with soil collected in the periphery, resulting, after a year, in the bleaching of the rock, obtaining better reading of the engravings for this. The continued application of this technique will allow a more rigorous survey without resorting to fungicides.

In 2009 we had written that the carapace was “made solely by blocks of Milky quartz of medium and small size.” a position that we now reviewed due to information obtained in the excavation of the 4th Tumulus of Vilares. Indeed, today we think that coverage would be exclusively made with large blocks of Milky room and some shale and greywacke, according to the abundance of each material in the site. It was verified, in this and in other counties around, that not all burial mounds are built with milky quartz, as there are, for example, only in quartzite pebbles (Abrantes and Pampilhosa da Serra), only in shale (Pampilhosa da Serra, Sertã, Arganil), and angular quartzite (Pampilhosa da Serra). Over the centuries this stone tends to fragment by the action of the elements, which induced many researchers, including us, to consider that was used a small stone layer to cover.

It is due to these factors that some engravings in the same rock were not drawn, as happened in the 19th Rock of Bregada that was only designed the figurative anthropomorph which has under it a contemporary boot drawing covered with lichens. It is necessary to cover with dirt the remaining engravings to get the whole composition.

We base our hypothesis in two observations made in tumulus. Firstly, the excavation in ditches revealed that the first layers of the carapace of the tumulus, consists of large blocks that have never been exposed to the atmospheric elements. Secondly, the excavation of the destryed chamber, revealed a filling of schist slabs (the cover of the chamber) and small quartz stones. If we imagine that the looting started at the top of the carapace, as would be natural, and where the stone was exposed to atmospheric elements, it is understandable that at the bottom of the chamber was only found very fragmented material.

2nd Rock of Vale do Gato The 2nd Rock of Vale do Gato, turned and tilted to the South, situated in an area that suffered afforestation, was almost entirely covered with dirt, that had to be removed so that can be drawn the engravings. The rock fully engraved. The inscriptions are from two left hands with part of the arm, recorded by percussion, dated by António Martinho Baptista of the Bronze age (figure 8.13), a little hole, only impacted, a tool or instrument made with a thinner perforation, which may indicate an earlier era or later, like a shovel, perforated clouds of difficult interpretation and a possible unfinished or destroyed semicircle.

As to its chronology, although written in 2009 that there were coals of 5th Tumulus of Vilares, these were so small and at such a small amount that it is very difficult to date.

99 

 

Figure 8.13: Drawing of 2nd Rock do Vale do Gato. 10th Rock of Bregada In the case of 10th Rock of Bregada we only drawed the main figure, because the rest of the panel has many lichens that difficult reading. The rock, with a medium leaning, is facing West and presents large amount of engravings.

Contains an engraved figurative drawing, that appears to be a shaman or upper entity, with antennas, high hands, seeming to hold an arm in his right hand, with big ears, engraved eyes and mouth and sexual attribute well demarcated, fitting in the Bronze age (figure 8.14).

  Figure 8.14: Partial drawing of Bregada’s 10th Rock.

100 

 

Figure 8.15: Total 11th Rock design Bregada. Rock XI de Bregada

21th Rock de Bregada

In the case of the 11th Rock it was necessary to remove some accumulated dirt in order to get the full extent of the rock. This presents one of the most important representations of rock art in the area and is found in a gully that receives the waters of a land path. The 11th Rock of Bregada contains a scene of pastoralism, with an anthropomorphic figure standing on what appears to be a cow with a rope (lead); at the foot seems to be represented also a dog with sexual attribute drawn. Ahead of the cow there is another animal that could be a goat, though it isn’t clear of what that is (Figure 15). The panel ends with the presence of a cruciform and also has a perfect little dimple.

This is a new engraved panel, facing South, which is on a dirt path and it was covered by land at the time of the making of Carta Arqueológica do Concelho de Pampilhosa da Serra, reason why we didn’t found it during the field survey. Because it is found on a forest path is likely to disappear with the floor leveling, frequent situation in forestry exploitation areas. The engravings are in a moderate condition, although they have suffered from the action of a leveling machine that destroyed a small part of the engravings. It features a full drawing of a foot and a circle surrounding an engraving inside (solar representation?) to which was added a dotted perhaps in an attempt to create a shoe or boot sole.

Below this rock, but very close, are other engravings, covered with lichens. It was tried the drawing of some of them, although the drawing may be subject to correction if, after the removal of the lichens, there is missing elements.

Relation between Rock Art and Tumuli

19th Rock of Bregada This rock panel is very large, not having been chances to fully draw this one. Although this panel is large and has a good surface for recording (cristalized surface), it hasn’t a large amount of engravings, and so the engravings are quite separated from each other. It is in a horizontal position, with the engravings facing the South. Between the old engravings arise, by trying to copy, some boot sole drawings recorded by contemporary pastors, as well as adding recent elements to the prehistoric engravings.

Rock art panels from the area of Bregada seem to be from the period of the Bronze Age burial mounds, like the 5th and 6th tumuli of Vilares located about 1200 m to the East. There are burial mounds even closer (about 200 m) with formal characteristics very similar to these, not yet interventioned. Engraved panels are between the heights of 960 and 835m. The burial mounds of Bregada are in the heights of 790 m and the ones from Vilares at 820 m. Both them have several rock engravings in their vicinity. By insufficient field studies we will not analyze in detail the various engravings, nor explain what they mean. We know that older art exists in this area, which can belong to the neo-chalcolithic period, but is not minimally 101 

 

studied. Although the majority of those who have already been analyzed and studied belong to the Bronze Age, some engravings appear to be oldest. Final Words The interventions carried out had an emergency nature; the monuments are, on the one hand very destroyed and, on the other, running serious risks to disappear completely. Apart from the excavation of the 4th tumulus of Vilares, the priority is for the excavation of monuments that are quite destroyed and in places that provide a rapid deterioration of the remains that still exist, as in the case of burial mounds that are located under forest paths and firebreaks and tend to disappear by action of heavy machinery. In the medium and long term, it is intended to carry out various studies about the burial mounds preserved and about their relationship with the identified rock art. In the field of rock art, much needs to be done: full excavation of the panels, which in many cases are partially covered with soil, lichen cleaning and the register of the engravings.

In the field of exploration of the rock art, there are still areas where burial mounds and rock art panels go unnoticed to our gaze. Since April 2009, public presentation date of Carta Arqueológica do Concelho de Pampilhosa da Serra, till today, it has been already identified six more burial mounds and several panels with rock art. The municipality is vast and mountainous and the steps that we have taken to find these traces were not sufficiently meticulous and to identify all existing traces. The approximately 150 tumuli already detected are scattered in an area of approximately 350 km2 and constitute the largest group of tumuli of the country, many of them in close association with about 150 rock art panels, within the municipality, since the engravings extend to the neighboring municipality of Covilhã. They are found in very rugged terrain that firstly were thought desert or little tempting to prehistorical populations, demonstrating an enormous vitality by pastoralist communities that made this area its permanent habitat and where they buried their dead. The study of geomorphology and landscape archaeology, could contribute, notably, to the understanding of why the territorial occupation of a very inhospitable area, where the cold and the wind are felt frequently and where snow is constant, in the cooler months.

References BATATA, C., (1998) – Carta Arqueológica do Concelho da Sertã. Sertã: Câmara Municipal de Sertã, 96 p. BATATA, C., (2008) – Prospecção Arqueológica no âmbito da Revisão do PDM de Amarante. Lisboa: DGPC – Direcção Geral do Património Cultural, final report. BATATA, C.; GASPAR, F., (1994) – Levantamento arqueológico do Concelho de Pampilhosa da Serra. Pampilhosa da Serra: Câmara Municipal de Pampilhosa da Serra, 64 p. BATATA, C.; GASPAR, F., (2000) – Arte Rupestre da bacia hidrográfica do rio Zêzere. In BUENORAMÍREZ, P.; CARDOSO, J.L.; DÍAS-ANDREU, M.; HURTADO, V.; JORGE, S.O.; JORGE, V.O. (coord.) – Pré-história Recente da Península Ibérica. Porto: ADECAP – Associação para o Desenvolvimento da Cooperação em Arqueologia Peninsular, actas do 3º Congresso de Arqueologia Peninsular, 4. BATATA, C.; GASPAR, F., (no prelo) – Pampilhosa da Serra, Património Arqueológico Revisitado - 15 anos depois. Pampilhosa da Serra: Câmara Municipal de Pampilhosa da Serra, actas das I Jornadas de História Local, 10 e 11 de Abril de 2008. BATATA, C.; GASPAR, F., (2009) – Levantamento arqueológico do Concelho de Pampilhosa da Serra. Pampilhosa da Serra: Câmara Municipal de Pampilhosa da Serra, 189 p. BATATA, C.; GASPAR, F., (2009a) – Escavação de emergência da Mamoa V de Vilares (Unhais-oVelho, Pampilhosa da Serra). Lisboa: DGPC – Direcção Geral do Património Cultural, final report.

BATATA, C.; GASPAR, F., (2009b) – Mamoas e arte rupestre no concelho da Pampilhosa da Serra (Centro de Portugal). AÇAFA On Line. Vila Velha do Ródão: AEAT – Associação de Estudos do Alto Tejo, 2, p. 2-21. BATATA, C.; BORGES, N.; CORREIA, H.; SOUSA, A. (2008) – Carta Arqueológica do Concelho de Vila Pouca de Aguiar. Vila Pouca de Aguiar: Câmara Municipal de Vila Pouca de Aguiar/OZECARUS, Serviços Arqueológicos, Lda, 191 p. CANINAS, J.; HENRIQUES, F.; BATATA, C.; BATISTA, Á., (2004) – Novos dados sobre a PréHistória Recente da Beira Interior Sul. Megalitismo e Arte Rupestre no Concelho de Oleiros. Coimbra: Separata da revista Estudos de Castelo Branco, Nova Série, nº 3. CANINAS, J.; SABROSA, A.; HENRIQUES, F.; MONTEIRO, J. L.; CARVALHO, E.; BATISTA, Á., CHAMBINO, M.; HENRIQUES, F. R., MONTEIRO, M.; CANHA, A.; CARVALHO, L.; GERMANO, A. (2008) – Tombs and rock carvings in the Serra Vermelha and Serra de Alvélos (Oleiros - Castelo Branco). In BUENO-RAMIREZ, P.; BARROSO-BERMEJO, R.; BALBÍNBERHMANN, R. (COORD.) – Graphical markers and Megalith Builders in the International Tagus, Iberian Peninsula. Oxford: BAR Publishing, BAR International Series 1765. CANINAS, J. C.; HENRIQUES, F.; BATISTA, Á.; CHAMBINO, M.; HENRIQUES, F. R.; MONTEIRO, M.; CANHA, A.; CARVALHO, L. (2009) – Estruturas monticulares antigas na fronteira 102 

sul do concelho do Sabugal. SABUCALE. Sabugal: Revista do Museu do Sabugal, nº 1. CRUZ, D. J. da (1990) – A Mamoa 2 de Chã de Carvalhal (Serra da Aboboreira, Baião). Trabalhos de Antropologia e Etnologia. Porto: SPAE, volume 31. CRUZ, D. J. da; et. al. (2000) – O grupo de Tumuli do Pousadão (Vila Nova de Paiva, Viseu). Estudos PréHistóricos. Viseu: Centro de Estudos da Beira Alta, volume VIII. DIAS, P. T. das N. (1985) – Subsídios para o levantamento arqueológico do concelho da Pampilhosa da Serra. Coimbra: trabalho apresentado à cadeira de Técnicas de Investigação Arqueológica da Faculdade de Letras da Universidade de Coimbra, manuscript. GASPAR, F. (2009) – Relatório preliminar da escavação de emergência na Mamoa do Cabeço da Linteira, Pampilhosa da Serra. Lisboa: DGPC – Direcção Geral do Património Cultural, final report. GASPAR, F.; BATATA, C. (2011) – Arte rupestre e monumentos funerários da pré-história recente do concelho de Pampilhosa da Serra. Report of the archaeological work progress carried out under a research project. HENRIQUES, F.; CANINAS, J.; CHAMBINO, M. (2008) – Carta Arqueológica de Vila Velha do Ródão – uma leitura actualizada dos dados da Préhistória Recente. In BUENO-RAMIREZ, P.; BARROSO-BERMEJO, R.; BALBÍNBERHMANN, R. (coord.) – Graphical markers and Megalith Builders in the International Tagus, Iberian Peninsula. Oxford: BAR Publishing, BAR International Series 1765. NUNES, J. de C. (1952) – Um machado de talão, de tipo

galaico, na Beira-Litoral Interior. Arganil: Museu da Câmara Municipal de Arganil. NUNES, J. de C. (1956) – O Ídolo pré-histórico das Relvas. Revista de Guimarães. Guimarães: Sociedade Martins Sarmento, LXVI, n.º 1/2. NUNES, J. de C.; PEREIRA, A. N. (1974) – A Pedra Riscada. Sá da Bandeira: separata da Revista dos Cursos de Letras, I. SILVA, F. A. P. da (1997) – Contextos funerários da Idade do Bronze nos planaltos centrais do CentroNorte Litoral português: tradição ou inovação? Neolítico, Calcolítico y Bronce. Zamora: Fundación Afonso Henriques, actas do II Congreso de Arqueología Peninsular, Tomo II. SANTOS, A. T.; BAPTISTA, A. M. (2011) – Rock art in the Iberian central chain: the cases of Piódão (Arganil) and Vide (Seia). In BUENO-RAMIREZ, P.; CERRILLO-CUENCA, E.; GONZALEZ CORDERO, A. (eds.) – From the Origins: The Prehistory of the Inner Tagus Region. Oxford: BAR Publishing, BAR International Series 2219, p. 161-176. SIMÕES, L. M. P. C. (s. d.) – Monografia da Pampilhosa da Serra. Manuscript. THADEU, D. (1951) – Geologia e jazidas de chumbo e zinco da Beira Baixa. Boletim da Sociedade Geológica de Portugal. Porto: Sociedade Geológica de Portugal, IX, fasc. I-II. VILAÇA, R.; SANTOS, A. T. M.; PORFÍRIO, E.; MARQUES, J. N.; CORREIA, M.; CANAS, N. (2000) – O povoamento do I Milénio a.C. na área do concelho do Fundão: pistas de aproximação ao seu conhecimento. Estudos Pré-Históricos. Viseu: Centro de Estudos Pré-históricos da Beira Alta, volume VIII.

103 

104 

 

Between norm and variation in the semiotic of the funerary world: examples and discussion of some abnormal graves in the Bronze Age Europe Davide Delfino

Abstract The study of funerary contexts can sometimes be flawed due to a monotonic and generalist approach to the contexts. Especially in rituals. But customs related to death may not sometimes be uniform, in the same region and at the same time being exceptions to “normalcy”, which points to the need to carry on studying funerary and ritual contexts using a more open and varied approach. In this paper we will present examples of “unusual” funerary rituals across Bronze Age for each region and each period in which they occur, which can cause problems if we base the interpretation of funerary contexts on a too processualist approach. We will present cases related to Italy, Serbia, Spain and Portugal, which are symptomatic of funerary rituals that deviate from normality of the typical archaeological record for some Bronze Age human cultures, but fit perfectly in the normal variation of a cultural and semiotic manifestation which sometimes is not uniform as is the case of funerary rituals. Keywords: Bronze Age, Europe, Burials, Symbolism, Semiotic, Abnormalities Introduction The interpretation of archaeological data is often based on a materialist/structuralist approach: the belief that human reason and action are structural (which is partly real), justified on many occasions by bringing the archaeological data together as if they were pieces of “dominoes” and structuring them would bring forward the various manifestations of ancient cultures. But this raises two problems: 1) often not all archaeological data can be structured and, in this case, some “dominoes” end up being censored because otherwise the “dominoes” do not match; 2) on the occasions where it is true that a researcher thinks and acts in a structuralist perspective, he/she often makes the mistake of thinking that the structure of our civilization, of our times, is the same as that of the ancient civilizations. This last problem, caused by the subjective structuralist approach, has already been raised by Criado Boado (2006: 250). Liable to overcome this convenient method that inserts and processes quickly as in a computer, the archaeological data seen from an extreme structuralist point of view need support from the cultural anthropology approach. But cultural anthropology does not submit to structuralist “religion”. When studying ancient societies one must take into account that human society acts according to a variable/invariable system, i.e. individual variation exists because their inherent differences intertwine with

common similarities. . Also the variation is not absolute but relates to the invariable and cultures themselves are made of this variable/invariable web (Remotti, 2011: 3435). In this sense, culture is subject to large but also small variations (ibid.: 89-107). Regarding the nature of these small deviations, they can be of hexogen or internal nature: the first is given by interactions with the outside and the second by a movement of internal spontaneous growth. A common factor is the reason to associate the two modalities of small social changes: the decision to adopt them as part of the whole human community. Can they be defined as part of the generality of these small changes? Yes, probably as anomalies with respect to the common standard. What is an anomaly? Literally it is a deviation from the norm. For the Anomalists of the school of Pergamon the irregularities in the formation and decline of the words are the fundamental principle of language (Perutelli, Paduano, Rossi, 2010: 16). In the context of human culture, not all abnormalities are the fruit of or synonymous with change, but have to do with a number of choices made to give character to some aspects of society. Despite skepticism among archaeologists, it is undeniable that, especially in funerary horizons, there are anomalies within burials of the same culture. In general terms, you have already abnormalities when, in a given structure, the individual units are no longer perfectly equal to each other. In this perspective, the abnormalities are present in many funerary horizons in a same cultural area starting, for example, with the big difference between the positions of the body according to gender. Such is the case with the position of the bodies and the composition of the grave goods at the Late Bronze Age Terramare culture (Northern Italy) in which anomalies between male and female are the norm that characterizes the way this culture interred their dead. Or, as the anomaly in different funerary architectures at Early Bronze Age of Cyclades islands is typical of this culture. But, among the abnormalities that are the norm there are two types: 1) General and with a low degree of anomaly (like the examples above); 2) In detail and with high index of anomaly (like the examples that will be provided in this paper). In this context, can the funerary horizons be defined? The author believes that they have, for the human community, a dual function which results in a dual nature: cultural and semiotic. Cultural Nature Purpose: preserve ancestral customs (and consequently the community customs) Semiotic Nature Purpose: communicate. The burial is a symbol to convey some kind of meaning. . It is relatively 105 

 

feasible to perceive the “what?” Generally the “two big messages” that you can understand from the various aspects (architecture, position of the grave, funerary treatment of the body, position of the body, number of dead in the same grave, grave goods) are: 1) Cultural affiliation and 2) Individuality of the dead. But, it is difficult to understand “to whom?” i.e. the targets of the funerary message. It is possible to speculate that it might be the gods from afterlife, death itself or posterity. But, we end up never knowing with absolute certainty. Talking about culture, we have to mention semiotics and signs: “Everything can be a sign” (Fidalgo & Gradim, 2004: 14) and “Culture is in all, from a certain point of view, the practice of unlimited semiotics” (Volli, 2005: 8). So, by virtue of its nature, also the semiotics of funerary horizons have to take into account the value of the anomalies in the construction of a language: “The language is not, therefore, a pure system of signs but a system of figures you can use to build the signs” (Saussure cit. by Volli, 2005: 14). “The easiest way to express what it means to a sign, is the sign to say what is not, what it is different, rather than trying to explain what” (ibid.: 15) Considering the fundamental role of irregularities in semiotics, it is considered important to focus on the anomalies in funerary rituals. But, we outline the anomalies of a second type (“In detail and with high index of anomaly” as described just above) since these constants in the cultures of the Bronze Age, and beyond, are the key towards understanding some of the parts of the message arising from funerary horizons. Choice of the samples areas The purpose of this paper is to show data and archaeological facts demonstrating that the presence of ritual burials that are considerable abnormal, when compared with the general overview of the archaeological material culture, are to be considered the norm, rather than the exception. In order to have a correct view of these anomalies that are the norm, tree levels of European geographical scale will be considered: 1) A macro region, which during Bronze Age had a similar social development dynamics verified over time and strong long-range interactions, mainly in metallurgical models. 2) Within each macro region, focus areas where, if possible, there are burial sites in sufficient number and relatively not too dispersed through a wide area. 3) Burials having, as much as possible, a very narrow chronological difference. Specifically to the first criteria, the choice goes to large areas of northern Italy, Balkans and the south of the Iberian Peninsula: during much of the Bronze Age these

tree major areas are known to have had similar development levels in social structures. At the Early Bronze Age the Polada Culture, in Northern Italy (2100-1600 BC), located in an alpine lakes region, and the first phase of Bronce de la Mancha, in central Spain (2000-1700 BC) defines a society not very complex yet, but by way of being (Gilman, Fernandes Posse, Martin 1996; Brodsky, Gilman, Martin Morales 2013: 142). In Middle and Late Bronze Age, we are going to witness the formation of increasingly pre-urban societies: Terramare in Northern Italy (1600-1150 BC) in the low Po valley and the second phase of Bronce de la Mancha in central Spain (1700-1500 BC); part of Vatin Culture in Western Serbia (1700-1200 BC), characterized by a chiefdom structure settlement and by its connection with Mycenaean area (Tásic 1996). For the Final Bronze Age, in the macro-regions represented by Polada and Terramare culture there is a partial collapse, especially in the evidence of settlements, but also, probably in the evolutionary continuity, in the parts of these areas which did not had an adequate policy management (Bernabó Brea, Cardarelli, Cremaschi, 1997; Cupitó, et. al., 2012) or have not stayed in a good position to take advantage by themselves of the major coastal (intra-Mediterranean) or inland (trans-European) traffic routes (Bietti Sestieri, 1997: 758-759). Also in the Atlantic façade of the Southwest Iberian Peninsula, the Bronze do Sudoeste defined by Schubart (1975) and recently revisited by Torres Ortiz (2008) and Oosterbeek, et. al. (2011) marks an important human cultural scenario, where the social pre-urban structure development was born in full Middle Bronze Age and completed in Final Bronze Age II with the beginning of Phoenician contacts. Data from Polada Culture Located in Center Northern Italy between 21st – 17th century BC (figure 9.1: area 1). This culture takes its name from the village on stilts of Polada. Many settlement types like this, commonly dated with ceramic and metallic artifacts related to 2100-1600 BC, characterize the Bronze Age period in the area between the Alpine lakes of Lombardy-Veneto and the Po River (De Marinis, 2000: 98-99). Generally, the settlements are characterized by villages on stilts, as Lavagnone di Desenzano, Fiavé, Polada, Lucone, installed in small lakes between moraines around some large alpine lakes in Northern Italy (Barich, 1971; Perini, 1994; De Marinis, 1997; De Marinis, 2000). Material culture joined with the fortunate combination of radiometric dating and stilt dendrochronology studies, from these villages, allows establishing a precise chronology, especially for the Early Bronze Age (table 9.1):

106 

 

Northern Italy Chronology E.B.A. Ia

Correspondence with Reinecke Chronology BronzezeitA1

Absolute Chronology

Material Culture

21st-start 20th cent BC

E.B.A.Ib

BronzezeitA2a

20th cent. BC

E.B.A. Ic

BronzezeitA2a

19th cent BC

E.B.A. II

BronzezeitA2b

end of BC

Polada pre Barche di Solferino pottery style/ Falhertz metallurgy Polada pre Barche di Solferino pottery style/ Bronze Metallurgy Polada pre Barche di Solferino pottery style/ Bronze Metallurgy Polada Barche di Solferino pottery style/ Bronze Metallurgy

19th- 17th cent

Table 9.1: Polada Culture chronology; Drawn from Reinecke, Bohner, Wagner, 1965; De Marinis, 1999). In the early chronological stages the links between the Polada area and the transalpine area are very important, passing on to the Adige (Trentino), showed for example in the Keulenkopfnadeln pin type (De Marinis, 2000: 108; 2003: 63). Unknown tombs or necropolis on most southern villages on stilts are to be found, but only in the northern part of Lake of Garda and in the middle Adige valley; the common characteristics are the fact that they are outside the settlements, in rock shelters and in lithic cists under small tumuli, adults in crouched or supine position and children in a pot. Again, the deposition is on the left or right side without distinction of sex (adults), often the

skull is removed and laid aside and in some cases there is double burial (De Marinis, 2003). Between the necropolis sufficiently known and related to Polada Culture, judging from the type of grave good objects (ceramics, metals or adornment elements in bone), there are many that can fit into the “pre Barche di Solferino” phase and then to Early Bronze Age I phase: Romagnano Loch, Mezzocorona Borgonuovo, Nogarole di Mezzolombardo, Catsel Corno di Mori, Santuario di Lassino and La Vela di Valbusa and Acquaviva di Besenello. All the burials in this necropolis, or singular grave, conform to the laws of normality between them, according to the characteristics mentioned just above, with exception of La Vela di Valbusa and Acquaviva di Besenello.

Figure 9.1: Localization of the archaeological Cultures treated in this paper: Polada Culture (1); La Mancha Bronze (2); Vatin Culture (3); Terramare Culture (4); Bronze Age: centre of Portugal (5).

107 

 

Figure 9.2: La Vela di Valbusa grave: context (over the smelting kiln), drawn from Fasani 1992. La Vela di Valbusa and Acquaviva di Besenello tombs La Vela di Valbusa is a tomb burial of an individual female, in an apparently lying position, even if it seems to be a secondary one, according to the rite of the double burial. It was placed under a small mound consisting of a base of large boulders to which stones of decreasing size have been overlapped. The structure was placed against the wall of a scarped rock, oriented NW / SE. Early Polada pottery and various elements of the personal ornament compose the grave goods (Fasani, 1990; Nicolis, 1997). Parallels with other tombs or necropolis are immediate, as for example, with Romagnano Loch (Perini, 1975): the same pottery, the same necklaces (same grave goods), the same position regarding one burial of Romagnano (tomb 13), the same ritual of the double burial in many graves, the same architecture (under tumuli) and the same landscape positioning. Also, parallels in burial ritual are found in Mezzocorona Borgonuovo and shelters 2 and 3 from Nogarole di Mezzolombardo (De Marinis, 2003: 18) The anomaly in the context of La Vela tumuli grave is the presence of a smelting kiln under the tumuli structure: in fact, after the removal of the burial, appeared a large area of several hundred copper smelting slag, mixed with carbon, with increasing intensity towards the rock wall coinciding with the mound and extending northward outside of it. In the center of the area covered by slags, is remarkable the presence of a baked clay quadrangular surface: it rested on a bed of slag with other slags encrusted on its surface, which was also deformed by the stones of the base of the burial mound. In the baked clay surrounding area were also found three bellows nozzles (tuyères) in ceramic (Fasani, 1990: 172, 174), (figure 9.2).

Similar to the tomb of La Vela di Valbusa is the one of Acquaviva di Besenello: it is another burial of a female individual, accompanied, as grave goods, by arrowheads and almond-shaped cusp, in association with a smelting copper kiln (Angelini, Bagolini, Pasquali, 1980). This, however, should be considered with extreme caution: the burial is prior to the implantation of the smelting kiln, unlike the other of La Vela, where the tomb is above the kiln and may be contemporary. The apparent anomaly of burials located on top of a smelting copper kiln, hides in reality another anomaly that raises an interesting question: if the metallurgical work was supposedly made by men (for obvious reasons of extreme conditions that required some physical prowess), why are two women buried in the kilns? Data from La Mancha Bronze South West Spain between 20th – 16th century BC (figure 9.1: area 2). It was defined by the Siret brothers in late 19th century (1890) as part of Argaric culture and is substantially different from the previous period in type of settlement, funerary ritual and material culture. A better and much recent definition distinguish this cultural area from Argar, defining it as Bronce de La Mancha (Nieto, Meseguer, 1988) or Bronce manchego. The area is defined at south by Segura river, at North by Tagus river, at Least the Vilanopó river. Currently we tend to characterize it as a distinct cultural horizon but with strong relationships with Argar and Valencian Bronze. In recent definition of the second phase of Bronce de La Mancha, we can observe a hierarchical system of settlements with large fortified settlements in the valley (motillas) and small settlement in hilltop (castellones). The former are understood as places of storage, monitoring of the crafts and subsistence goods while the 108 

 

La Mancha chronology Early Bronze

Correspondence with Reinecke Chronology BronzezeitA1BronzezeitA2a BronzezeitA2BronzezeitB1

Absolute Chronology

Social evidences

Non-structured settlement system. Middle Bronze 18th - 16th cent. BC Ever more hierarchical society, with few male warriors armed with swords and many burials of children. Structured settlement system. Table 9.2: La Mancha Bronze chronology; Drawn from Reinecke, Bohner, Wagner, 1965; Nieto, Meseguer 1988 and Castro, et. al., 2001: 181. latter are seen as having been used for territorial control. In economy, agriculture is the main function, followed by farming and metallurgy, still not bronze, but the arsenical copper. Material culture is typologically represented by riveted daggers and undecorated carenated pottery (Brodsky, Gilman, Martin Morales, 2013: 141). In general it is not very different from Argar Culture. This, with the recent research, particularly Castro, et. al. (1993) is a clearer perspective considering the aspects in common with Bronce de La Mancha: - hilltop settlements are distinguishable but not uniform in the way of settlement; - burials inside the settlements or the houses constitutes a radical novelty regarding the Chalcolithic, and its uniqueness within West Europe Early Bronze Age; but not all funerary manifestations fall within this dominant criteria. In the aspect of chronology, in the last 20 years, it was possible to collect a sufficient number of samples for radiometric dating, which allowed data to be combined with the relative chronology of metals. Nevertheless it does not seem to have shed much light on the various phases of argaric occupation – la Mancha and Valencian Bronze settlements and necropolis. It is possible, however, to systematize it in a table (table 9.2). As for the burial context, it seems that during all Bronce de La Mancha, the mode of burial was maintained substantially equal, like in Argar culture. Despite what has been La Mancha Early and Middle Bronze Age two very wide chronological phases, it is normal to find substantial variation in the funerary ritual. This shows two differences in the grave goods between motillas (simple) and castellones settlement type (much rich and varied), but uniformity exists within ritual according to the age: adults in lithic cist, children in pottery vessel (in most cases) (Meseguer, Galan, 2004).

21st - 18th I cent. BC

Concerning these general “common low” in burial, there are some exceptions, like La Motilla del Azuer. La Motilla del Azuer (Ciudad Real) (figure 9.3) Intervened since 1974 by T. Najera and F. Molina (Granada University), this fortified settlement had very big defensive walls that conferred the site with a fundamental importance inside the Argar Culture. The fortified settlement is in the plain and the radiometric date shows an occupation between 2450 and 1540 BC (Najera Colino, et. al., 2006: 151). Fortifications are composed by a central quadrangular tower, a first wall around the tower, and a second external and cyclopic (with very big stone) wall. The village is outside this second wall. Inside the fortification, were found 39 graves, of adult or children, characterized by: singular grave, in simple pit or lithic cist, children in urn (ibid.: 153-154). Grave goods are poor. Interesting is grave nº 39: near by the external wall façade, with several reconstruction stages along time, associable to the settlement phase 5, dating from 1.800 BC. Architecture grave consists in a lithic cist, within which a male child was buried (8-9 age old) in flexed position and lateral decubitus. Grave goods consist of pottery miniaturist material: 4 vessels, 2 clay disk and 1 toy clay; and also, one stone ball (Ibid.: 153-154). The singularity of this burial is: 1) children not inhumed in urn, but as adults in lithic cist; 2) position of the grave towards the fortification wall; 3) miniaturist pottery in grave good (also with very poor technology, like in the bonfire). As miniaturist pottery, there are also other examples in Argar area in various burials, or domestic contexts (Sanchez Romero, 2004; Aranda Jimenez & Molina Gonzalez, 2005). As for the other two singularities, worth of note is the position of the grave inside the fortification wall, in connection with the fortification, and the absence of the typical urn within which it was usual to bury children.

109 

 

Figure 9.3: La Motilla del Azuer. Grave n. º 39: the position in the motilla (1); the grave (2); plan of the grave (3). Drawn from Najera Colino, et. al., 2006. Vatin chronology Early (Pančevo-Omoljica) Middle (Vatin-Vršac) Late (Belegiš I-Ilandža)

Reinecke chronology Bronzezeit A2-B1 Bronzezeit B2-C Bronzezeit C-D

Absolute chronology 20th – starting 15th cent. BC 15th – cent. BC 14th – starting 12th cent. BC

Table 9.3: Vatin Culture chronology; Drawn from Garašanin 1983. Data from the Vatin Culture The Middle Bronze Age in Serbia is chronologically set in the 15th century BC (figure 9.1: area 3). Bronze Age in Serbia is still insufficiently known. Uneven geographical studies caused Bronze Age period to look like a patchwork formed of many cultural groups. The cultural relationships among the groups, their chronology and territorial settlement would have to be more accurately defined (Lijustina, Dimitrovic, 2009: 53). Generally it is characterized by the Vatin Culture with a tripartite generic chronology between the 16th and the 13th centuries BC (Garašanin, 1983), (table 9.3): This culture suffered some influences of Mycenaean Greece and had already developed social differentiation. Also, it has developed large central settlements, which were surrounded by smaller settlements and farms (Tásic, 1996). As for funerary horizons, the personal prestige status emerging during the Middle Bronze Age pursuits a new burial fashion, born in Early Bronze Age. Funerary practices occasionally become uniform over broad areas, although burials are remarkable both for their regional and chronological diversity. The practice of single graves under tumuli is widespread during the first half of Bronze

Age. Variety of tumuli and good graves suggest that access to status, power and wealth had been prerogative of restrict social groups. Grave architecture is characterized by tumuli, with inhumation in Early phase and in general incineration in Middle phase (Lijustina, Dimitrovic, 2012: 37), with both rituals during the beginning of Middle phase (Lijustina, Dimitrovic, 2013: 109). Dubac necropolis (Jančići) Mound III, grave 5 The beginning of Middle Bronze Age necropolis of Dubac consists in three tumuli made of earth and stones. In each tumulus we have several graves, mixing incineration (in full Middle Bronze Age) and inhumation left in crouched position turned right (early Middle Bronze Age). From a total of 2 graves, grave 5 of mound III is the most interesting. It is an individual inhumed female over 50, placed at the center of mound III and having three interesting data: 1) it is datable to the final phase of the Middle Bronze Age (differing from the other graves in this phase, with incineration); 2) it is centrally placed in this mound, reserved for distinct or elderly person (in mound I and II the central grave is of males); 3) it is positioned unusually, stretched on the stomach (ibid.: 109-110), (figure 9.4: number 1).

110 

 

Data from the Terramare Culture At eastern Terramare from Late Bronze Age with a 13th century BC chronology (figure 9.1: area 4). It is located south of Po river and in the Venethian plain formed by big villages on stilts surrounded by embankments. Fully developed in the Middle Bronze Age 2 (1500-1450 BC), it involves a massive colonization of the Po Valley, with intensification of agriculture and material culture caracterized by pottery with handles, horned superelevation and connection with italian peninsula (protoAppennic area). After a period of implementation and prosperity during the Middle Bronze Age 3 (1450-1300 BC), in the Late Bronze Age, population pressure increases with greater social complexity. Metallurgical production reaches its maximum level, alongside with an increase in fortification of the largest villages. In the production of ceramics there is a clear break with the western Po Valley area. In this period the eastern Terramare are in contact with traffic and people coming from the Aegean through the Adriatic Sea: Frattesina and other settlements in the Po’ river delta have emporiums with Mycenaean pottery and Baltic Amber processing. At the beginning of the Late Bronze Age, the Terramare undergo a sudden collapse, with the exception of eastern villages (Veneto). Probably caused by a missed ability to control the depletion of the natural resources caused by the intensive use of soils. Veneto, probably a more evolved social

structure, aided by a greater cultural dynamism across the sea and the Alps, escaped from sudden collapse (Bernabó Brea, Cardarelli, Cremaschi, 1997; Bettelli, Cupitó, 2010; Cupitó, et. al., 2012). The graves in the Olmo di Nogara and Franzine Nuove necropolis In the Terramare context, in low valleys near Verona, are necropolises from Middle Bronze Age 2 ( Olmo di Nogara) , Middle Bronze Age 3 (Bovolone and Scalvinetto) and Late Bronze Age ( Franzine Nuove and Vallona di Ostiglia) with both rituals, inhumation/incineration, and positioned at the top of a small hill near the settlement (Salzani, 2005). Inhumed individuals are in lying position; a minority of male have a warrior grave good; minority of femals have rich personal adornments (Leonardi, 2010: 28). In the Olmo di Nogara and Franzine Nuove necropolis, we have two inhumation graves, of individuals lying on the stomac. In the Olmo di Nogara, it is an adult female individual (grave 56), (Roncoroni, 2006: 67). In the Franzine Nuove, it is a male young-adult individual, whose head is turned to left and lower limbs are flexed (Corrain, Capitanio, Fasani, 1967). Apparentely doesn’t have grave goods (figure 9.4: number 2).

Figure 9.4: inhumed laid on the stomach: Dubac (Middle Bronze Age of western Serbia); Franzine Nuove (Late Bronze Age in eastern Terramare). Drawn from Lijustina, Dimitrovic, 2013 and Corrain, Capitanio, Fasani, 1967. 111 

 

Southwest Bronze III Chronology

Correspondence with Reinecke Chronology

Absolute Chronology

Final Bronze Age I

Bronzezeit D/ Hallstatt A1/A2

12th cent. BC

Evidences

Begin of really bronze metallurgy Huelva phase Middle 9th- Middle Burnished decoration Hallstatt B1 10th cent BC pottery/Atlantic metallurgy Baiões/Plaza de las Middle 10th- end 9th Burnished decoration Monjas phase Hallstatt B2 cent. BC pottery/Atlantic and Mediterranean metallurgy Colonial phase Post end 9th cent. BC Phoenician influxes, Hallstatt B3/ hole pottery and iron Hallstatt C1 metallurgy Table 9.4: Southwest Bronze III chronology; Drawn from Reinecke, Bohner, Wagner 1965; Torres Ortiz 2008 and Oosterbeeck, et. al., 2011. Bronze Age: centre of Portugal Final Bronze Age in Central Portugal dates between the 12th – 18th century BC) (figure 9.1: area 5). Territory includes Tagus, Zêzere and Nabão rivers hydrographic basin, with southern part of Beira Interior. It is, currently characterized by a not very well defined archaeological culture for the Bronze Age, very dependent of relative chronology and material culture studies in Baiões/Santa Lúzia, on the north (Senna Martinez, 2010) and Bronze do Sudoeste, on the south (Schubart, 1975; Parreira, 1995; Cardoso, 2007), what can today be considered as a cultural appendix. A much specific subdivision was made possible for the Bronze do Sudoeste III (Southwest Bronze III), by comparing with Southwestern Spain (Huelva) material culture (specifically the metallurgy), settlements modalities and chronologies. So, we will use the Bronze do Sudoeste as a chronological reference (table 9.4): The material culture is represented by smoothed decorations in a pottery (Lapa do Fumo – Alpiarça type) particularly reticule decoration, Atlantic metallurgy (double ring axes, tongue carp swords) and after influxes by Mediterranean (fibulae Pantalica type, for example); in these phases hilltop settlement in hinterland and familiar farms near the rivers are the predominant characteristics. All the data show a society in transition to a more structured level: starting in Chalcolithic (Oosterbeeck, et. al., 2011: 149, 151) it develops with the tin trade in the Final Bronze Age (Senna Martinez, 2013) and continued with the arrival of the Phoenician, at the begining of the Iron Age, along the Tagus, directly until Santarém, indirectly for the higher territorial area (Arruda, 2005; Delfino, 2012).

About the funerary horizons, three fundamental characteristics are highlighted in the North Alentejo – Tagus Valley area: - Replacement of the ancient monumentality of the graves by much more discrete architecture like mamoas (flat tumuli) like Souto (Cruz, 2011) or burials in lapas (little caves) like Gruta da Marmota (Kunst, 1995: 124), or in necropolises like Urns Fields, like Tranchoal (Cruz, Vilaça, Gonçalves, 1999), Meijão (Kalb, Hock, 1985), and Cabeço da Bruxa (Kalb and Hock, 1982), or necropolises in cist, like Provença and Pessegueiro (Kunst, 1995, 124). - Funerary ritual is mixt with burial (grave in lapas) and incineration (“field” necropolis and mamoas) - Continuity of funerary usage of the necropolis since Neolithic until Bronze Age, like Gruta do Morgado Superior (Cruz, et. al., 2013) Monte de São Domingos (Castelo Branco) (figure 9.5) Worked by J.C. Caninas in 1997 (Cardoso, Caninas, Henriques, 1998) this is probably a settlement constituted by two circular structures, with the bottom composed by an upright flat stone. Structure 1 is the smallest and shows a pavement in clay with charcoals and other evidences of everyday life (like three fragmented pottery shapes). Structure 2 is the biggest, doesn’t have evidences of everyday life, but under the pavement three little quartz lumps had been built that cover three pits. One of the pits contains one pottery urn with cremation (burned bones and ashes). Anomalies in the funerary world by the end of Final Bronze Age, in Central Portugal are: 1) graves in a domestic context; 2) relative monumentalization of the graves; 3) incineration graves in domestic context and in a relative monumental structure. Although it is also reasonable to understand structure 2 as a specifically funerary structure (doesn’t have material associated with daily life and is bigger than structure 1), the anomaly of a grave inside the settlement area remains. 112 

 

Figure 9.5: Monte de São Domingos: structure 1 (1); incineration urn in structure 1 (2); burned bones of urn (3). Drawn from Cardoso, Caninas, Henriques, 1998. Discussing Before some considerations, it is important to present some reflections made in the last 30 years that probably already pointed the interpretation methodology towards a certain tendency: “Postulate according to which if the data could be put into a computer, would split in a convenient taxonomy that would reveal cultural groups, is called into question” (Renfrew, 1987). “A semiotic approach in the study of cultures is crucial: Material manifestation = “Parole”; Idea producing material = “Langue”; Anomalies must be considered with the same attention of the customs; Culture = set of signs; Sign = set of differences that distinguish it from other signs” (resumed from Tabaczynsky 1996). “Nevertheless it is quite frequent to be tented to understand these aspects (funerary horizons) in constant, pursuing statically and unidirectional codes for each cultural and geographical situation” (Roncoroni, 2006: 63) Also from the work of F. Roncoroni (ibid.) it is possible to understand that customs in funeral rituals can return

centuries later and in adjacent regions. This demonstrates that rituals do not always follow systematic rules in prehistoric society but depend on semiotic laws. Conclusions The problem about the message that the authors of the burials wanted to convey is twofold: What did they want to communicate and to whom? For the “what they wanted to communicate”, the answers can be manyfold and, most of the times, none of them is really confirmable: the overall response “wanted to communicate the life of men and women”, or rather the actions, the social affiliation, the consideration that the community and the family had for them, perhaps it is the only one that can be taken as certain, because of the clear evidence material. Major problems faced by questions are: Who you want to allocate these messages (posterity, the afterlife gods, death itself?). Often, the only material evidence (place of burial, architecture, location of the deceased, cinerary ritual, good grave) is not sufficient to provide clear indisputable explanations. . Does “stomach inhumation” such as in Dubac, Olmo di Nogara and Franzine Nuove have a particular meaning? If we infer the meaning of similar burials in historical 113 

 

times, we come across individuals who had been excluded from the life of the community when they were alive: witches, people considered frenzied, people considered affected by vampirism. Or simply were guilty of serious crimes. Can we envisage a similar motivation for the Bronze Age according to a general axiom relating to a pre-existing knowledge (Delfino, Charters de Almeida, 2013: 125-126) or, on the contrary, is there a completely different cultural context from the European historical epoch, and so, almost impossible to verify these days? Finally, in the context of a particular common rule, in general, the burials within a human community are, by contrast, subject to standardization, specification of the message within the funerary horizon. An important aim to remember: one chrono-cultural phase (for example Early References ANGELINI, B.; BAGOLINI, B.; PASQUALI, T. (1980) – Acquaviva di Besenello (Trento). Preistoria Alpina. Trento: Museo Tridentino di Scienze Naturali, nº 16, p. 67-69. ARANDA JIMENEZ, G.; MOLINA GONZALEZ, F. (2005) – Intervenciones arqueologicas en el jacimiento de la Edad del Bronce del Cerro de la Encina (Monachil, Granada). Trabajos de Prehistoria. Madrid: Consejo Superior de Investigaciones Científicas, volume 62, nº 1, p. 165180. ARANDA JIMENEZ, G.; ESQUIVEL, J.A. (2006) – Ritual funerario y comensalidad wen las sociedades de la Edad del Bronce del Sureste peninsular: la Cultura de El Argar. Trabajos de Prehistoria. Madrid: Consejo Superior de Investigaciones Científicas, volume 63, nº 2, p. 117-133. ARANDA JIMENEZ, G.; MONTON SUBIAS, S.; SANCHEZ ROMERO, M.; ALARCON, E. (2009) – Death and everyday life. The Aragaric society from Southeast Iberia. Journal of Social Archaeology. [S.l.]: Sage journals, volume 9, nº 2, p. 139-162. ARRUDA, A.M. (2005) – O Iº milénio a.n.e. no Centro e no Sul de Portugal: leituras no início de um novo século. O Arqueólogo Português. Lisboa: Museu Nacional de Arqueologia, série IV, volume 23, p. 9156. BARICH, B. (1971) – Il complesso industriale della stazione di Polada alla luce di piú recenti dati. Bullettino di Paletnologia Italiana. Roma : Museo preistorico etnografico L. Pigorini, 80, p. 77-182. BERNABÓ BREA, A.M.; CARDARELLI, A.; CREMASCHI, M. (1997) – Il crollo del sistema terramaricolo. In BERNABÓ BREA, A. M.; CARDARELLI, A.; CREMASCHI, M. (eds) – Le Terramare: la piú antica civiltá padana. Milano: Elekta, catalogo della mostra di Modena, p. 745-756. BETTELLI, M.; CUPITÓ, M. (2010) – Fando Paviani, Legnago (Verona). In RADINA, F.; RECCHIA, G. (eds) – Ambra per Agamennone. Indigeni tra Micenei tra l’ Adriatico Ionio ed Egeo. Bari: Adda Editore, catalogo della mostra di Bari, p. 258-259.

Bronze Ib in Polada Culture, or Huelva phase in Southwestern Iberia Peninsula) covers about hundred years or several decades and during this relative long period, traditions may be maintained or on the other hand new trends, ideas or other factors transforming the society are likely to appear. And, these transformations, including the funerary ritual, also depend on personal and familiar sentiments or the overall sentiments of the small community. There are exceptions within the standard. These should not be underestimated as an integral part and physical evidence of an ancient language the community wanted to transmit. And which perhaps, as archaeologists, we are involuntary receivers. And sometimes the abnormalities of the language are the key to understanding where normality is lacking. BIETTI SESTIERI, A.M. (1997) – Il territorio padano dopo le Terramare. In BERNABÓ BREA, A. M.; CARDARELLI, A.; CREMASCHI, M. (eds) – Le Terramare: la piú antica civiltá padana. Milano: Elekta, catalogo della mostra di Modena, p. 757-769. BLANCE , B. (1971) – Die Anfange der metallurgie auf der Iberische Halbinsel. Studien zu den Anfangen der Metalurgie. Berlin: Mann, nº 4. BRODSKY, M.; GILMAN, A .; MARTIN MORALES, C. (2013) – Bronze Age political landscape in La Mancha. In CRUZ BERROCAL, M.; GARCIA SANJUAN, L.; GILMAN, A. – The Prehistory of Iberia. Debating Early social stratification and the State. New York: Routledge, p. 141-169. CARDOSO, J.L. (2007) – Pré-história de Portugal. Lisboa: Verbo. CARDOSO, J.L.; CANINAS, J.C.; HENRIQUES, F. (1998) – Duas cabanas circulares da Idade do Bronze do Monte de São Domingos (Malpica do TejoCastelo Branco). Estudos Pré-Históricos. Viseu: Centro de Estudos Pré-Históricos da Beira Alta, p. 325-345. CASTRO, P. V.; COLOMER, E.; CHAPMAN, R.; GILI, S.; GONZALEZ, P.; LULL, V.; MICÓ, S.; MONTÓN, S.; PICAZO, M.; RIHUETE, C.; RISCH, R.; RUIZ PARRA, M.; SANAHUJA, M. E.; TENAS, M. (1993) – Projecto Gatas. Sociedad y economia en el sureste de Espanha c.2.500-800 antes de nuestra era. Huelva: Junta de Andaluzia, I.A.A., 1985-1992 Projectos, p. 401-415. CASTRO, P. V.; CHAPMAN, R.; GILI, S.; LULL, V.; MICÓ, S.; RIHUETE, C.; RISCH, R.; SANAHUJA, M.E. (1998) – Aguas Project. Paleoclimatic reconstruction and the dynamics of human settlements and land-use in the area of tha Middle Aguas (Almeria) in the South east of the Iberian Peninsula. Luxemburgo: European Commission Directorate XII. Science, Research and Developpment. CORRAIN, C.; CAPITANIO, M.; FASANI, L. (n.d.) – Un inumato in posizione bocconi nella Necropoli Enea di Franzine Nuove di Villabartolomea (Verona). Quaderni di Antropologia e di Etnologia. Nº 4, p. 145-148. 114 

 

CRIADO BOADO, F. (2006) – Se puede evitar la trampa de la subjectividad? Sobre arqueoologia e interpretación. Complutum. Madrid: Universidad Complutense Madrid, nº 17, p. 247-253. CRUZ, D.; VILAÇA, R.; GONÇALVES, A. (1999) – A Necrópole de Tanchoal dos Patudos (Alpiarça, Santarém). Conímbriga. Coimbra: Instituto de Arqueologia, nº 38, p. 5-29. CRUZ, A. (2011) – A Pré-História Recente do vale do baixo Zêzere. Arkeos, perspectivas em diálogo. Tomar: Centro Europeu de Investigação da PréHistória do Alto Ribatejo, volume 30, PhD dissertation, Universidade de Trás-os-Montes e Alto Douro, 273 p. CRUZ, A.; GRAÇA, A.; OOSTERBEEK, L.; ALMEIDA, F.; DELFINO, D. (2013) – Gruta do Morgado Superior – Um Estudo de Caso Funerário no Alto Ribatejo (Tomar, Portugal). Las Crisis en La Historia: Noción y Realidades. Comunidad de Castilla-La Mancha: Departamento de la Universidad de Castilla – La Mancha, Vinculos de Historia, nº 2, URL: vinculosdehistoria.com/índex.php/vínculos/article/vi ew/62, p. 143-168. CUPITÒ M., DALLA LONGA E., DONADEL V., LEONARDI G. (2012) – Resistances to the 12th century BC crisis in the Veneto region: the case studies of Fondo Paviani and Montebello Vicentino. In KNEISEL, K..; KIRLEIS, W.; DAL CORSO M.; TAYLOR, N.; TIEDTKE V. (eds) – Collapse or continuity? Environment and development of Bronze Age human landscapes. Proceedings of the international workshop Socio-Environmental Dynamics over the Last 12,000 Years: The Creation of Landscapes, II, Kiel, 14th-18th March 2011, 1, p. 55-83. DE MARINIS, R.C. (1997) – L’etá del Bronzo nella regione Benacense e nella Pianura Padana a Nord del Po’. In BERNABÓ BREA, A. M.; CARDARELLI, A.; CREMASCHI, M. (eds) – Le Terramare: la piú antica civiltá padana. Milano: Elekta, catalogo della mostra di Modena, p. 405-422. DE MARINIS, R.C. (1999) – Toward a relative and Absolute Chronology of the Bronze Age in Northern Italy. Notizie Archeologiche Bergomensi. Bergamo: Civico Museo Archeologico di Bergamo, nº 7, p. 23100. DE MARINIS, R.C. (2000) – Il Museo Civico Archeologico “Giovanni Rambotti” di Desenzano del Garda; un’ introduzione alla preistoria del Lago di Garda. Desenzano del Garda: Cittá di Desenzano del Garda. DELFINO, D.; CHARTERS DE ALMEIDA, J. (2013) – Uma nova abordagem para um diálogo entre arqueologia e arte. A Vénus pré-histórica da Colecção Estrada, In JANA, I.; PORTOCARRERO, G.; DELFINO, D. (eds) – Actas II e II Jornadas Internacionais do M.I.A.A. Museu Ibérico de Arqueologia e Arte. Abrantes: Câmara Municipal de Abrantes, p. 117-128. FASANI, L. (1990) – La sepoltura e il forno di fusione

della Vela di Valbusa (Trento). Preistoria Alpina. Trento: Museo Tridentino di Scienze Naturali, nº 11, p. 165-181. FIDALGO, A.; GRADIM, A. (2004) – Manual de Semiótica. Covilhã: Universidade de Beira Interior. GARASANIN, M. (1983) – Zapadnosrpska varijanta vatinske grupe. Praistorija jugoslavenskih zemalja. Sarajevo: Bronzano doba, nº IV, p. 736-753. GILMAN GUILLEN, A.; FERNANDEZ POSSE, M. D.; MARTIN, C. (1996) – Consideraciones cronológicas sobre la Edad del Bronce en La Mancha. Complutum. Madrid: Universidad Complutense Madrid, nº 6, extra II, p. 111-137. GONZALEZ MARCEN, P. (1994) – Cronologia del Grupo Argarico. Revista de Arqueologia de Ponent. Lleida: University of Lleida History Department, nº 4, p. 7-46. KALB, P.; HÖCK, M. (1982) – Cabeço da Bruxa, Alpiarça (Distrito de Santarém) – relatório preliminar da escavação de Janeiro e Fevereiro de 1979. Portugália. Porto: Universidade do Porto. Faculdade de Letras, Nova Série, volume II/III, p. 61-69. KALB, P.; HÖCK, M. (1981–1982) – Cabeço da Bruxa, Alpiarça (distrito de Santarém). Relatório preliminar da escavação de Janeiro e Fevereiro de 1979. Portugalia. Porto: [s.n.], Nova Série, nº 2–3, p. 61– 69. KALB, P.; HÖCK, M. (1985) – Cerâmica de Alpiarça. Exposição Temporária na Galeria dos Patudos Catálogo. Alpiarça: Câmara Municipal de Alpiarça/Casa-Museu dos Patudos/Instituto Arqueológico Alemão de Lisboa. KUNST, M. (1995) – A Idade do Bronze na Estremadura. In OLIVEIRA JORGE, S. (ed.) – Idade do Bronze em Portugal. Discursos de poder. Lisboa: Instituto Português dos Museus, catálogo da exposição do Museu Nacional de Arqueologia, p. 124-125. LEONARDI, G. (2010) – Premesse sociali e culturali alla formazione dei centri protourbani del Veneto. Bollettino di Archeologia on line I. Rome: Proceeding of International Congress of Classical Archaeology; meeting between cultures in the ancient Mediterranean (Rome 2008), volume speciale F/F1/4, www.archeologia.beniculturali.it/pages/pubblicazion i.html, p. 23-35. LJIUSTINA, M.; DIMITROVIC, K. (2009) – The Bronze Age Vatin Culture in the west Moravia Basin- case study of Sokolica in Ostra. In ZANOCI, A.; ARNAUT, T.; BAT, M. (eds) – Studia Archaeoligiae et Historia Antiquae. Doctissimo viro Scientiarum Archeologiae et Historiae Ion Niculiţă, anno septuagesimo aetatis suae, dedicator. Chişinău: Univ. de Stat din Moldova, p. 53-63. LJIUSTINA, M.; DIMITROVIC, K. (2012) – In search for prestige: bronze age tumular graves in West Serbia. In SIRBU, V.; SCHUSTER, C. (eds) – Tumuli graves. Status symbol of the death in the Bronze and Iron Age in Europe. Oxford: BAR Publishing, Proceedings of the XVI U.I.S.P.P. World Congress in Florianopolis (4-10 September 2011), BAR Int. Ser. 2396, p. 35-42. 115 

LJIUSTINA, M.; DIMITROVIC, K. (2013) – Middle Bronze Age in tumular graves in Kablar range, West Serbia- Gender perceived as physical condition and social construct. In SIRBU, V.; MATTEI, S. (eds) – Bronze and Iron Age graves from Eurasia - Gender between Archaeology and Anthropology. Buzau: Muzeul Judetean Buzau, Proceeding of the 13th Internacional Colloquium of Funerary Archaeology (Buzau 17th- 21th October 2012), Mousaios, XVIII, p. 105-129. LULL, V. (2000) – Argaric society: death at home. Durham: Durham University, volume 74, nº 281, p. 581-590. MESEGUER, J. L.; GALAN, C. (2004) – El Serro de La Encantada. In GARCIA HUELTA, M. R.; MORALES HERVAS, J. (eds) – La Penisnula Iberica en el II Milennio a.C.: Poblados y Fortificaciones. Cuenca: Ediciones de la Universidad de Castilla-La Mancha, p. 115-172. NAJERA COLINO, T.; MOLNA GONZALEZ, F.; SANCHEZ ROMERO, M.; ARANDA JIMENEZ, J. (2006) – Un enterramiento infantil singular en el yacimiento de la Edad el Bronce de La motilla del Azuer (Daimel- Ciudad Real). Trabajos de Prehistoria. Madrid: Consejo Superior de Investigaciones Científicas, volume 63, nº 1, p. 149156. NICOLIS, F. (1997) – La sepoltura dell’ Antica Etá del Bronzo della Vela di Valbusa. In ENDRIZZI, L.; MARZATICO, F. (eds) – Gli ori delle Alpi, catalogo della mostra di Trento. Trento: Provincia Autonoma di Trento, p. 372-373. NIETO GALLO, G.; SANCHEZ MESEGUER, J. (1998) – Bases para la sistematización del estudio de la Edad del Bronce de La Mancha. Pueblos y Culturas Prehistóricas y Protohistóricas. Castilla-la-Mancha: Servício de Publicaciones de la Junta de Comunidades de Castilla-la-Mancha, Iº Congreso de Historia de Castilla-La Mancha, volume 2, p. 221227. OOSTERBEECK, L.; DELFINO, D.; BAPTISTA PEREIRA, F.; PORTOCARRERO, G. (2011) – Objectos, economias e sociedade nas origens da urbanização no Sudoeste peninsular. Museu Ibérico de Arqueologia e Arte de Abrantes, antevisão III. Abrantes: Câmara Municipal de Abrantes, p. 149154. PARREIRA, R. (1995) – Aspectos da Idade do Bronze no Alentejo Interior. In OLIVEIRA JORGE, S. (ed.) – Idade do Bronze em Portugal. Discursos de poder. Lisboa: Instituto Português dos Museus, catálogo da exposição do Museu Nacional de Arqueologia, p. 131-134. PERINI, R. (1975) – La necropoli di Romagnano Loch III e IV – Le tombe all’inizio dell’ Etá del Bronzo nella regione alpina SudOccidentale. Preistoria Alpina. Trento: Museo Tridentino di Scienze Naturali, nº 11, p. 295-315. PERINI, R. (1994) – Scavi archeologici nella zona palafitticola di Fiavé - Carena. Trento: Servizio Beni Culturali della Provincia Autonoma, volume 3. PERUTELLI, A.; PADUANO, G.; ROSSI, E. (2010) –

Storia e testi della letteratura latina. Bologna: Zanichelli. REINECKE, P.; BOHNER, K.; WAGNER, F. (1965) – Mainser Aufsatze zur chronologie der Bronzer und Eisenzeit. Bonn: Habelt ed. REMOTTI, F. 2011) – Cultura dalla complessitá all’impoverimento. Bari: Laterza editore. RENFREW, C. (1987) – Problem in the modelling of socio-cultural system. European Journal of Operational Research. London: Elsevier, volume 30, issue 2, p. 179-192. RONCORONI, F. (2006) – Riflessioni sull’interpretazione dei riti funerari in arqueologia preisttorica e proistorica. Bulletin d’Études Prehistoriques et Archeologiques Alpines. Aoste: Société Valdôtaine de Préhistoire et d’Archéologie, nº XVII, p. 63-69. SANCHEZ ROMERO, M. (2004) – Children in the Southeast of Iberian Peninsula during Bronze Age. Ethnographish-Archaeologische Zeitschrift. Leipzig: Universität Leipzig, nº 45, p. 377- 387. SALZANI, L. (2005) – La Necropoli dell’età del Bronzo all’Olmo di Nogara. Memorie del Museo Civico di Storia Naturale di Verona. Verona: Museo Civico di Storia Naturale, Sezione Scienze Dell’uomo, Serie 2, nº 8. SENNA MARTINEZ, J.C. (2010) – Un mundo entre mundos. O Grupo Baiões/Santa Luzia: sociedade, metalurgia e relações inter-regionais. Iberografias. Guarda: Centro de Estudos Ibéricos, nº 6, p. 13-26. SENNA MARTINEZ, J.C. (2013) – Um rio na (s) rota (s) do estanho: o Tejo entre Idade do Bronze e Idade do Ferro. CIRA Arqueologia. Vila Franca de Xira: Museu Municipal de Vila Franca de Xira, nº 2, p. 718. SHUBART, H (1975) – Die kulture der Bronzezeit in Sudwestern der Hioberischen Halbinsel. Berlin: Walter de Gruyter, 2 volumes. SIRET, E.; SIRET, L. (1890) – Las primeras Edades del metal en sudeste de Espanha. Barcelona. Edición facsimilar, 2006, Murcia: Museo Arqueológico de Murcia. SPINDLER, K.; CASTELLO-BRANCO, A.; ZBYSZEWSKY, G.; FERREIRA, O. Da VEIGA (1973-1974) – Le monument a coupule de l’Age du Bronze Final de la Roça do Casal do Meio (Calhariz). Lisboa: Serviços Geológicos de Portugal, Comunicações do Serviço Geológico de Portugal, nº 57, p. 91-153. TABACZYNSKI, S. (1996) – La cultura ed i suoi corrrelati oggettuali; approccio pragmático, approccio semiótico, approccio integrale. Ferrara: A.B.A.C.O. Edizioni, Atti del XIII Congresso U.I.S.P.P. di Forlí- Italia (8-14 Settembre 1996), volume 1, sottosezione: problemi teorici e metodologici, p. 3-7. TÁSIC, N. (1996), (ed.) – The Danube Yugoslavian Basin and the neighboring regions in the 2nd millennium BC. Vrsac-Belgrade: Institute for Balkan Studies, proceeding of International Colloquium of Vrsac 1995. TORRES ORTIZ, M. (2008) – The chronology of the 116 

 

Late Bronze in Western Iberia and the beginning of U.I.S.P.P. Congress, Lisbon September 2006, British Archaeological Reports International Series 1871, p. the Phoenician colonization in the Western 135-148. Mediterranean. In BRANDHERM, D.; TRACHSEL, M. (eds.) – A new dawn for the Dark Age? Shifting VOLLI, U. (2005) – Manuale di Semiotica. Bari-Roma: Laterza. Paradigms in Mediterranean Iron Age chronology. Oxford: BAR Publishing, proceedings of the XV

117 

118 

 

The polimorphism of graves and the distribution of archeological remains in the Southwest Bronze Age necropolis of Soalheironas (Alcoutim) João Luís Cardoso, Alexandra Gradim

1/25 000, folha 567, edição de 1991): 7º 30’ 00’’ long. W; 37º 30’34’’ lat. N.

Abstract The Southwest Bronze Age necropolis Soalheironas (Alcoutim)21 was excavated in 2005 by a team directed by the first author and with ongoing participation of the second, with the results already published in their essential form.

The site is located in a direct line about 0, 5 km to the west of the river Guadiana on their right bank which is deeply jagged by valleys of numerous tributaries (Figure 10.1).

The remarkable number of graves, more than thirty, which confirm this necropolis to be one of the most important of Southwest Bronze Age, allowed a careful consideration of their respective architecture and structural characteristics, which led to their typology based on the evidence of their diversity. The distribution of the archeological remains in the excavated graves was also subject of analysis and discussion. Resumo A necrópole do Bronze do Sudoeste de Soalheironas (Alcoutim) foi escavada em 2005, por uma equipa dirigida pelo primeiro dos signatários, com o apoio permanente do segundo e encontra-se já publicada, nos seus traços essenciais. A assinalável quantidade de sepulturas que a integram, mais de uma trintena, que fazem dela, uma das mais importantes do Bronze do Sudoeste, possibilitou uma observação detalhada da respetiva arquitetura e características estruturais, que conduziu à constituição de uma tipologia, tendo presente a diversidade evidenciada. A distribuição dos espólios arqueológicos pelas sepulturas escavadas foi também objeto de análise e discussão. 1 – Introduction. Geographic geomorphologic characteristics

location,

The Southwest Bronze Age necropolis of Soalheironas is located in eastern Alto Algarve (Council of Alcoutim), at the following geographic coordinates, taken from the Military Map of Portugal in the scale of 1/25 000 (sheet 567, edition 1991) (Carta Militar de Portugal à escala de

                                                             21

This work was written by the first author, based on records collected in the field by both authors. The first version of this English version was provided by the second author who wishes to thank Ute Bonillas for her help in providing the translation. The photos are of the first author and the figures were made by Bernardo Ferreira, coordinated by the first author based on field records of the Technical Bureau of Tavira, the second author and Fernando Estêvão Dias (Alcoutim Municipal Council).

Figure 10.1: Localization of the site in the Iberian Peninsula (above) and its implantation on a narrow and elongated, rocky ridge, 3 km from the left bank of the Guadiana river, in second plan. The necropolis is established in the terrain clearly delimited by its topography: it sits on a narrow and elongated, rocky ridge, which, with a height of 133 m at its crest, consists of regularly alternating beds of schist and greywacke of the Upper Carboniferous (“flysch” facies). Situated along this ridge, it reaches, well defined by the alignment of the graves, a length of more than 100m (Figure 10.2). In one previous paper where published the main results of the excavation of the entire site that was carried out in 2005, including the description 119 

 

Figure 10.2: Spatial distribution of the graves by types and correlated archeological artifacts. of the remains found (Cardoso & Gradim, 2008a, 2011). At the time, the remarkable diversity of the 32 excavated graves had not been adequately appreciated and therefore, this paper is preceding publication of a more rigorous study of this outstanding pre-historic necropolis, which is one of the most important among the known necropolises of the Bronze Age in the southwestern peninsula. In fact, necropolises with 13 or more graves constitute only 5% of all necropolises studied until 1998, with the majority (about 67%) having no more than 5 graves (Garcia Sanjuán, 1998). It is certainly one of the most interesting ones due to the close relationship between the distribution of the graves and the morphology of the terrain that was chosen for its peculiar characteristics.

The architectonic study, carried out on 32 excavated graves, confirmed a canonical orientation W/E, characterized by a rectangular pit, dug into the geologic substrate not deeper than 0, 50 m. The graves differ in the variable number of stone slabs that were used to demarcate each one of the cavities. In this paper are presented the results that were obtained from a typology ascribed to each tomb according to its location in the necropolis, and also their respective archeological remains. The objective is to verify or not verify any special relationship between the distribution in the terrain of better structured graves and their respective contends, in order to detect lineages that stand out from the community buried there.

Figure 10.3: Typology of the graves. 120 

 

2 – Classification of the graves According to the investigation carried out, four major types were identified, as shown in figure 10.3. However, it is important to consider these results with some precaution. While, in fact, it is unquestionable that tombs exist with between four slabs and none at all, it is also true that in some cases the severe disturbance of the surface of the terrain by heavy machinery during the recent widening of the forest road that runs along the necropolis, could have caused partial destruction of some tombs by crushing the tops of various slabs, or even total destruction of others. Type 1 – classic graves, commonly known as cists, which are of rectangular shape, defined by four slabs (Figure 10.4, above, on the left); Type 2 – tombs defined by three slabs, with one consistently missing on one of the longer sides. While in some cases the missing slab can be explained by the partial destruction of the tomb, in other cases it is evident that such slab never existed: as demonstrated in grave 29 (Figure 10.4, above, on the right), which was constructed by making use of the geologic outcrop, in this case consisting of a layer of greywacke sloping toward the interior of the grave, for one of the longer sides, making the placing of the respective slab unnecessary. Type 3 – tombs defined by two slabs, forming nearly always an angle in the structure (Figure 10.4, in the center, on the left). Indeed, of the six graves defined by

two slabs, only in one (grave 1) it was found that the two slabs corresponded to the two longer sides of the grave. Type 4 – tombs consisting of only one slab, placed nearly always on one of the longer sides of the grave. Of the four graves that incorporate only one slab, in eight cases such slab was positioned on the opposite side to the layer of greywacke sloping toward the interior of the grave, constituting the natural double, the reason why it was dispensed, as seen in other more elaborate graves of Types 2 and 3 (Figure 10.4, in the center, on the right). Type 5 – tombs that are simple pits dug into the geologic substrate and not defined by any slab, represented by only one, grave 32, the last that was excavated in the necropolis, owing to the difficulty of its identification (Figure 10.4, at the bottom). According to the above description it can be concluded that the distribution of the graves along the top of the ridge also respected the geologic structure of the outcrop. They were consistently oriented in a NW-SE direction that corresponds to the direction of the above mentioned alternated layers of schist and greywacke, sloping toward SW. It is evident that the placement of the tombs was determined by the existence of the more prominent layers of greywacke, which in the simpler constructions (graves of Types 2, 3 and 4), made up nearly always one of its longer sides. This is one of the more significant examples of constraints imposed by the geologic structure on the organization and construction of a pre-historic necropolis.

Figure 10.4: Examples of the five types of graves. Type 1 (above, on the left); Type 2 (above, on the right); Type 3 (in the center, on the left); Type 4 (in the center, on the right); Type 5 (at the bottom). 121 

 

3 – Relationship between the typology of the tombs and their spatial distribution

4 – Relationship between the classification of the graves and their respective remains

One of the more interesting facts concerning the polymorphism of the graves, represented and identified by the five above types, resides in the eventual spatial relationship that exists between the graves of each type. In Figure 10.2 stood out the various sections in which the necropolis is organized, based on the natural grouping of the graves that are part of it. It would be therefore important to ascertain up to which point such groups are correlated with the typology of the respective graves.

As it is common in this type of necropolis, the remains that were collected were scarce and of distinct importance and significance. Their distribution throughout the 32 graves in this necropolis is very interesting, considering their distinct concentration in the eastern part of the necropolis. Figure 10.2 is explanatory in this respect as far as the origin of 2 copper artifacts, without doubt the most valuable elements found, as well as the ceramic offerings are concerned. In fact, while the 10 graves in the eastern part of the necropolis contained 10 ceramic vessels and the above mentioned copper artifacts, the remaining 23 graves that occupy the middle and the western part of the necropolis contained only 3 vessels.

Considering the grouping consisting of sectors 5, 6 e 7, which occupy the center of the necropolis as a single sector that totals eight graves, two of them have four slabs, four have three slabs and only two have one slab. On the other hand, sector 8, which also contains eight graves, has an identical distribution, with 2 graves of type 1, four of type 3 and two of type 4. The two easternmost sectors (sector 1 and 2) can be grouped together and add up to six graves, of which two have four slabs; finally, the other grave built with four slabs is located in the adjacent sector, sector 3. According to this assessment of distribution, it can be concluded that there is no clear concentration of more elaborate graves in any particular sector of the necropolis that would point to the existence of lineages with different social status. However, such conclusion is no indication of a non-existence of family lineages, as are suggested by the concentration of graves into sectors along the nearly 100 m length that the necropolis occupies. Therefore, wanting to associate a more elaborate grave with a more distinguished individual, this difference would not be expressed at the level of the whole community, but only at the level of a family group of which the occupant was the most outstanding member. This detail is, as a matter of fact, in direct parallel to the facts established in the more important and better classified necropolises of the period. In fact, in the necropolis of Atalaia (Ourique), the first one to be adequately investigated, the graves were found in a honeycomb arrangement surrounding a more important grave that belonged to the founder of each group in the necropolis (Schubart, 1975). The same situation was identified in the necropolis of Alfarrobeira (Silves): a single cluster of graves was found that spread out from an older and more important grave, belonging to a person with more prestige than the others who were buried around him (Gomes, 1994). This situation was also more recently found in the Monument II of Pessegueiro, an important necropolis in the region of Sines, which consists of 27 graves, all of them cists, grouped into four main clusters, each one of them using as a focal point one main grave (Silva & Soares, 2009). Without doubt, because of the spatial organization of the necropolis, it is the most similar to the one here described.

5- Discussion According to the above it is concluded that from the architectonic point of view there are no differences in the distribution of the five types of graves that were identified in the excavation. If a social hierarchy in this community existed, it was not expressed in the architecture of the graves, but yes, by the differentiated distribution of the funerary remains, which don’t appear to be homogeneous, as shown in Figure 10.2. There are various explanations possible. For example, the first one corresponds to the more ancient graves that occupy the elevated part of the necropolis, which contain more remains: once the upper part of the necropolis was fully occupied, it later extended to the gentle slope to the west. Consequently, it would be permissible to think of a community that became impoverished or, in alternative, the abandonment with time of the tradition accompanying their dead with offerings. In this explanation, the differing distribution of the funerary offerings doesn’t imply the existence of any social differentiation into the community. Other explanation for the concentration of the funerary offerings in the upper part of the necropolis accepts the existence of the above mentioned social differentiation, pointing to the more outstanding members of the community in the dominant sector of the grave yard. However, against this alternative is the evidence coming from the more important and better known necropolises of the same period in the south of Portugal. They were organized in a different way, by which the most important graves that stood out were not located in a specific location of the necropolis but were special in the sense that they were the focal point in respect to others that were later constructed around them. This pattern, however, is not what was observed in the majority of the Algarvian excavations of Bronze Age necropolis carried out during the end of the 19th century by Estacio da Veiga, and during the 20th century by various other archeologists (Gomes, 1994: 64) The absence of a hierarchy, demonstrated by the simple distribution of isolated graves in the field, point to an 122 

 

absence or a rudimentary social differentiation in each community.

recovered from the grave of Belmeque (Serpa) (Schubart, 1975).

Indeed, it is common to consider the Southwest Bronze Age as a culture in which a social hierarchy based on economy is seldom evident (Cardoso, 2007: 459). This is even more apparent in this arid region, inhabited by poor populations with few resources, except for the ones resulting from the mining of rare occurrences of copper that appear throughout the western Algarve. Estacio da Veiga had already correctly pointed out this fact for the western part of Algarve (Veiga 1891: 82). In this case, the exploration of copper resources was helped by the proximity of the river Guadiana, which was an important means for transport of products and people across the region. In these poor areas, each necropolis was associated with a settlement of small dimensions that blended into the landscape and were therefore difficult to identify in the terrain.

6- Conclusions

In this context it is important to consider, with due reservations however, the current situation in the region during the I Iron Age. In this period, the necropolises appear to express the existence of a clear social differentiation. The excavation of the necropolis of Cabeço da Vaca (Alcoutim), brought to light this evidence, where a cist of large dimensions belonging to a warrior whose outstanding social status justified his situation in a culminant point of the ridge, separated from the rest of the necropolis, situated in a lower plan of the elevation (Cardoso & Gradim, 2006, 2008 b). The spatial distinction of the archeological remains of the graves of Soalheironas, in which the richest grave, belonging to the most elaborate type of architecture is located in the richest part of the necropolis and containing two vessels and the single copper dagger found, can be thought of expressing the social status of the interred individual. It can be therefore inferred to be a distant predecessor of the subsequent reality that is observed in the same region during the I Iron Age. However tenuous, and dictated largely by the scarcity of disposable resources, a rough sketch of a social hierarchy already existed in the midst of this community. In actual fact, other regions of the Southwest Bronze Age would assume during this period clearer and unmistakable contours: it suffices to cite the rich funerary remains

1- The Southwest Bronze Age necropolis of Soalheironas with 32 individual graves presents itself as one of the most important funerary sites of this Culture that are found throughout the south of Portugal and part of the adjacent Spanish territory. Its location, deliberately established along a narrow and elongated ridge of Paleozoic rocks (schist and greywacke layers, flysch facies) and amounting to a length of about 100 m, represents a rare example of geomorphologic conditions imposed on the geometry of the necropolis and on the orientation of the graves. 2- In fact, the structural conditions of the massif with the stratification of the layers of greywacke sloping toward the NW quadrant, determined not only the orientation of the major axe alignment of the graves toward NE-SW, but also allowed in many cases the dispensation of the slab on the SE side, occupied by the layer of the greywacke with identical function. 3- The significant number of graves in this necropolis permitted their adequate architectonic classification, which led to the identification of five different types, however integrated in the same canonical model. Their spatial distribution revealed, nevertheless, remarkable uniformity, a sign that if there existed any social distinction in the community, it was not expressed in the quality of the architecture of the graves. 4- The funerary offerings were concentrated in the ten graves in the easternmost part of the necropolis that occupies the highest spot on the long, rocky ridge where it was established. Although various explanations may be possible, it can be admitted that there existed a beginning, however incipient, of social distinction in the community in agreement with the scarce resources available in the region that was inhabited by small settlements of poor people. This supposition is supported by the existence of one cist containing a set of grave goods, with parallels with others in different geographic areas of the Southwest Bronze Age that precede the evidence found locally in the necropolises of the I Iron Age, where social distinction is already clearly evident. e dos resultados obtidos. Promontoria. Faro: Universidade do Algarve, nº 6, p. 223-248.

References CARDOSO, J. L. (2007) – Pré-História de Portugal. Lisboa: Universidade Aberta. CARDOSO, J. L.; GRADIM, A. (2006) – A necrópole da I Idade do Ferro de Cabeço da Vaca I (Alcoutim). Silves: Câmara Municipal de Silves, Actas 3.º Encontro de Arqueologia do Algarve (Silves, 2005), nº 1, p. 201-226. CARDOSO, J. L.; GRADIM, A. (2008 a) – A necrópole de cistas da Idade do Bronze das Soalheironas (Alcoutim). Primeira notícia dos trabalhos realizados

CARDOSO, J. L.; GRADIM, A. (2008 b) – O núcleo II da necrópole da Idade do Ferro de Cabeço da Vaca (Alcoutim). Silves: Câmara Municipal de Silves, Actas 5.º Encontro de Arqueologia do Algarve (Silves, 2007), nº 1, p. 103-116. CARDOSO, J. L.; GRADIM, A. (2011) – Dez anos de trabalhos arqueológicos em Alcoutim. Do Neolítico ao Romano. Lisboa: Câmara Municipal de Alcoutim. GARCIA SANJUÁN, L. (1998), (ed.) – La Traviesa.

123 

 

Ritual funerário y jerarquización social en una comunidad de la Edad del Bronce de Sierra Morena occidental. Sevilla: Universidad de Sevilla/Exc.mo Ay.to de Almadén de la Plata, SPAL Monografías, nº 1. GOMES, M. V. (1994) – A necrópole de Alfarrobeira (S. Bartolomeu de Messines) e a Idade do Bronze no concelho de Silves. Silves: Câmara Municipal de Silves, Xelb, nº 2. SCHUBART, H. (1975) – Die Kultur der Bronzeseit im

Südwesten der Iberischen Halbinsel. Berlin: Walter de Gruyter & Co. SILVA, C. T.; SOARES, J. (2009) – Práticas funerárias no Bronze Pleno do litoral alentejano: o Monumento II do Pessegueiro. Estudos Arqueológicos de Oeiras. Oeiras: Câmara Municipal, nº 17, p. 389-420. VEIGA, S. P. M. Estácio da (1891) – Antiguidades Monumentaes do Algarve. Lisboa: Imprensa Nacional, nº 4.

124 

 

The faces of death: from the Bronze to the Iron Age, between the North and the South of the Portuguese territory Raquel Vilaça

On the trail of death The funerary practices of the late Bronze Age, such as of Iron Age, are not well known in the Central Western and Northwestern regions of the Iberian Peninsula, as well as generally, in its Atlantic façade. So begins a text which 14 years ago assessed and discussed the then available information from Portuguese central territory and neighbouring areas (Vilaça & Cruz, 1999). This synthesis, perhaps insufficiently disclosed, but yet still a constant quoting source in several papers, was part of a project22 which produced new data of which most are published and others are about to be published. Despite the new contributions, whether arising from the project itself or from the studies conducted by other researchers, the general lines put forward in that text have not undergone deep alterations with respect to the Central Region. Two main conclusions were then presented: on the one hand, and in contrast with the then dominant discourses, it was shown that there was mortuary material evidences with their own spaces, although not always easy to identify; and on the other hand, that there was the need to recognise diversity, i.e. not regularity, within funeral practices and rites of the late Bronze Age (Vilaça & Cruz, 1999: 84). Therefore, if any norm is perceptible, it is precisely the absence of it, either by the variability of spaces, structures, materials and rituals coexisting in time although not in the same spaces - with inhumations and incinerations, or by a tendency towards a certain invisibility through which death does (not) express itself, what by definition leaves room for multiple hypotheses. In face of the identification of new burial structures and the increasing value given to previously known situations, it was also sought to demonstrate that the assumed absence or rarity of burial spaces in the late Bronze Age of the Atlantic façade was rather a result of directly introduced general interpretations from extra peninsular Atlantic façade (a tendency for water deposition and incineration practices, for example), than actually from the almost unknown, although present, Portuguese empirical evidences. Moreover, the important

                                                             22

This investigation project, entitled “Práticas Funerárias e/ou Cultuais dos Finais da Idade do Bronze na Beira Alta” (“Funerary and/or Cultic Practices from Late Bronze Age in Beira Alta”), was approved by the Portuguese Archaeological Institute in 1998, continuing another, started in 1993. Coordinated, the latter by Domingos Cruz and the former by the authors of this paper, developed within CEPBA (Centro de Estudos Pré-Históricos da Beira Alta / Centre for Prehistoric Studies of Beira Alta), with a considerable amount of participants, among researchers and students, linked to several institutions.

works developed by Philine Kalb and Martin Höck in Viseu area had already refuted the claim that there were no burials from this period in Portugal (Kalb & Höck, 1979). That synthesis also emphasised that, besides the need for systematic research in this field, the analysis could not restrict itself to a classical perspective in which death was understood as being confined to formalised inhumations and incinerations, of burials in spaces designed or reused for this purpose and properly accompanied by the socalled “grave goods” (Vilaça, 1999: 180; Vilaça & Cruz, 1999: 76). Death and its rituals go far beyond. It should be noted that it is also crucial to view the death phenomenon in those archaic societies not as a moment, but as a process developed in several actions translated through funerary ceremonies, with distinctive times, scenarios and players. In reality, human existence involves three (and not two) phases: life, death and the transition from one to the other. Burial structures and funerary spaces are only part, and not necessarily the most important one, of a chain of passage rituals probably composing several scenarios: some, materialised in actions, which sometimes have left traces, others bodily, or based in bodily actions (postures, songs, prayers, dances, gestures, looks, etc.), vanished in the very moment, becoming lost memories that barely fit into Archaeology. This is the first challenge presented by the regional and chronologic research study of death. It is an “evasive”, inexpressive and often invisible death occasionally dispersed and certainly not always protected and ultimately with multiple facets and disguises whose track, when present, translates into different and even unexpected situations. And it is also the first certainty: the communities reacted with diverse answers to the unavoidable power of death. The second challenge, from which it depends, is obviously the absolute need to further develop and multiply research projects focused on these problems as available well-characterised data and contexts are not sufficient and merely suggest some clues that should be further explored. However, the abominable situation of the world known to us, and that we feel very close in our daily life, greatly limits and restricts the creation and materialisation of ambitious, time-consuming, expensive projects. Meanwhile, it is necessary to go through other paths, trying to explore the available information. Recently, in another paper (Vilaça, in press), we have revised longstanding findings and contexts of quite different interest because some only provide concise 125 

 

information,23 while others, such as Paranho incineration necropolis (Tondela) (Coelho, 1925; Cruz, et al., 1997), provide rather interesting information. Let us focus on other issues, starting with the prevalence of traditions and the resistance to change. Marks of ancestry Among the various actions and reactions to death by the late Bronze Age communities in the space and time under analysis are those who translate a link with the past, or if you like, a continuity of the past in (their) present, highlighting traditions and roots that were naturally reinterpreted, filtered and shaped by time (to use the Marguerite Yourcenar’s familiar metaphor). The most significant manifestation of this phenomenon is the symbolic and physical appropriation of ancient spaces, places and contexts, in other words, pre-existing structures that we may systematise in at least three types of situations. The first corresponds to cave reutilisation as long-lasting funerary spaces. Although with several problems still to be clarified (Vilaça, in press), it is likely that certain communities continued to bury their relatives in caves as their ancestors had done before within a certain line of continuity that seems to be underlying. The second includes situations where old structures, with or without alterations, are reused. In this set we may include the reuse of megalithic monuments with all the symbolic power that they always had across time – phenomenon present in several areas and whose investigation has seen, in the last few years, a renewed interest (v.g. García Sanjuan, 2005 and several subsequent texts, among other authors). In Beira Alta, one of them is the Rapadouro 1 (Pendilhe), monument from which rectangular chamber some human bones were gathered with traces of having been submitted to fire (Cruz, 2001: 111-113; 185). Another is monument 2 of Fonte da Malga (Côta, Viseu), featuring a simple chamber, in which tumulus a cist was inserted corresponding to a secondary burial dated from late Bronze Age (Kalb & Höck, 1979; Kalb, 1994). Other possible traces of the re-utilisation of ancient megalithic monuments correspond to findings of dispersed materials, such as those in Orca do Picoto do Vasco (Pendilhe) (Cruz, 2001: 186 and fig. 74-3), which may point to mere visits without an effective reuse, i.e. formalised, either in the Late Bronze Age or in the Iron Age.

                                                             23

It is the case of the apparent association of inhumation burials to metallic artifacts as registered in Casal de Santo Amaro (Sintra) and Vendas das Figueiras (Penela), as well as the cultic-funerary utilisation of some caves from the wide karstic spots developed in the coastal area of the western Atlantic façade of the Iberian Peninsula.

If we move from Beira onto Portuguese Estremadura, we are confronted with a more complex situation, where the funerary structures resemble tholos with Late Bronze Age materials. Whether these correspond to reuses of Calcolithic spaces or are rather directly articulated with its construction, is a pertinent question. This issue, discussed in due time (Vilaça & Cruz, 1999: 80) was worthy of attention among the case study revisions, as happened with Roça do Casal do Meio (Sesimbra) (Cardoso, 1999-2000: 400-405; 2000b) which we had the opportunity to be commented on (Vilaça and Cunha, 2005). The re-examination of the monument within the on-going project, “Valorisation of Archaeological Heritage as part of the application: Arrábida to World Heritage”24 should bring new contributions, namely related to their construction chronology and ritualisation of the monument’s surrounding areas. The person in charge of the excavation had admitted it might have been an indigenous structure of tholos type reused in the Late Bronze Age (Spindler, et al., 1973-74: 117-118), - hypothesis that would later be valorised (Belén, et al., 1991: 237). The issue is not, however, of unequivocal reading, because it also referred the collection, at the monument’s construction level, of ceramic fragments matching a carinated bowl, typical of the Late Bronze Age (Spindler, et al., 1973-74: 124 and 149). But it is important to note that on the assumption that it has been reused to house the two inhumed men around the 10th century BC, it must also be assumed that a complete “past cleaning” of the chamber contents has also been done, of which there are no remains (or perhaps any hasn’t survived) (Vilaça & Cunha, 2005). The third situation includes cases where a new construction is made, especially small tumuli, but in areas adjacent to or in the vicinity of the old monuments which are not directly reused as sacred places. In this case, the process is of a distinct nature, in which the same sites are changed and take on new, more complex scenarios, while maintaining the ancestral symbolism of the place. Paraphrasing David Fontijn, we may say that “mounds attracted mounds” (Fontijn, 2007: 73). This seems to have been the strategy of diverse human groups, as it was possible to ascertain in some Beira Alta contexts. One of the more emblematic and thoroughly studied cases is in the Sr.ª da Ouvida (Castro Daire) extensive plateau. About three dozens of small tumuli of different chronology (between about 1450/1400 and 800 BC) and also mounds of bigger size and volume embody a true “sacred field” (Cruz & Vilaça, 1999). Five of those monuments have been digged up and no cist- or pit-like structures have been found; instead there were slabs, blocks and bedrock natural depressions defining central areas. The clear presence, although minimal, of burned

                                                             24

We have collaborated in the project led by the Association of Setúbal Municipalities as scientific consultants of the Palimpsesto/Arqueohoje Consortium for archaeological and conservation/restoration works.

126 

 

Figure 11.1: General view from Sr.ª da Ouvida with the location of tumuli 13, 12, 11 and 10; hermitage in background (according to Cruz & Vilaça). wood remains under some tumuli allowed hypothesising that it would belong to remains of pyre incinerations carried out in the vicinity. Another example can be observed in the necropolis of the eight monuments in Fonte da Malga (Viseu), two of which are megalithic. There is a double situation here: a cist in the tumulus of monument 2 alongside new nonmegalithic monuments such as monument 1, with a diameter of 6m, which had a small central cist surrounded by cairn bounded by a stone circle, also assigned to the

Late Bronze Age, based on the ceramic fragment collected there (Kalb & Höck, 1979). Therefore, concerning its past, certain communities from final Bronze Age also expressed themselves in death in variable ways. Some have conferred and reinforced the importance of the symbolic meaning of old places with distinct solutions, bringing multiple temporalities together. These are places whose temporality is in essence “a temporality of sequence” (Lucas, 2005: 39).

Figure 11.2: Monument 1 of Fonte da Malga after excavation (according Kalb and Höck). 127 

 

Figure 11.3: Signaling of Tumuli 1, 2, 3 and 4 of Pousadão, in W/SW – E/NE direction (according to Cruz, et. al.). Gravestone Landscapes Small tumuli like those we have just mentioned may also occur, actually more frequently, in areas where no traces of dolmens (or elder structures) are found, as is observed, for example, in the four monuments in Pousadão (Vila Nova de Paiva) (Cruz, et al., 2000). Thus, and in parallel with the phenomenon described in the previous section, other occurrences like this one, are materialised by the creation of new clusters, that suggests a parallel dynamic, at least from late Middle Bronze Age to early Late Bronze Age, now guided by the appropriation of new territories. Undoubtedly, “the Bronze Age communities did not limit themselves to the re-utilisation of old tombs”. As Domingos Cruz noted, “they also have their own solutions consisting in building non monumental tumuli” (Cruz, 2001: 266). In summary, the phenomenon is characterised as follows: small tumuli standing on ridges and platforms overlooking the valley with reduced diameters and volumes, i.e. with low impact on the topographical landscape, but marked by the bichromatic contrast from the recurrent use of quartz pebbles as tumuli covers, which certainly also had high symbolic power. In some cases, looking for increased spatial marking, people seem to have privileged the proximity to natural bedrocks, as occurred in Pousadão, settled along a significant series of granite outcrops in a symbiosis where culture and nature mix together; the structures and contents are highly variable, with pits, cists, cist structures, central areas defined by pebbles, ashes, charcoal, vases and vase fragments, fire traces, etc., This variability is sometimes present in the same set, which may indicate complex

rituals shared among several buildings with different functions, specifically related to funeral and other ritual activities in the context of practices associated with death (Cruz and Vilaça, 1999: 159; Cruz, et al., 2000). This final aspect is important as it seems to reflect actions and gestures, perhaps complementary, towards different places of the same necropolis or funerary-ritual complex. The case of the Casinha Derribada tumuli group is particularly suggestive in this respect (Cruz, et al., 1998). Hence, tumuli are not built up to house the dead. They are empty tombs, cenotaphs, without bodies or even traces of their remains, inhumed or cremated. But they would have served as ceremonial spaces connected with those. The communities invest in death memorials but not explicitly in its preservation. The chronology is set for some cases: we face a constructive phenomenon that may date back to the Early Bronze, as Serra da Muna (Viseu) monument 2 case (Cruz, et al., 1998 a), but dates mainly from the end of the Middle Bronze or Late Bronze Age, as revealed by monument 3 of Casinha Derribada (Viseu) (XV-XII centuries BC) (Cruz, et al., 1998 b) and Senhora da Ouvida 7 (Castro Daire) (XIV-XII centuries BC) (Cruz & Vilaça, 1999). However, the question of chronology is an aspect that deserves further research, admitting the possibility that some may have persisted to the Iron Age (Santos & Marques, 2007: 40), or even to historical periods (Middle Age). Actually, similar constructions from other peninsular areas such as some of the over 1100 registered in the Pyrenean region (Peñalver, 2005: 302) also fit in

128 

 

Chart 11.1: Graphics dates: Carbon 14 dates related with funerary contexts from Central Portugal late Bronze Age. (“Radiocarbon Calibration Program rev. 6.1.0”, Stuiver, M. et Reiner, P.J., 1998, Radiocarbon 35, 215-230; Reimer et al., 2009). PJ Reimer, MGL Baillie, E Bard, A Bayliss, JW Beck, PG Blackwell, # C Bronk Ramsey, CE Buck, GS Burr, RL Edwards, M Friedrich, PM Grootes, # TP Guilderson, I Hajdas, TJ Heaton, AG Hogg, KA Hughen, KF Kaiser, B Kromer, # FG McCormac, SW Manning, RW Reimer, DA Richards, JR Southon, S Talamo, # CSM Turney, J van der Plicht, CE Weyhenmeyer (2009) Radiocarbon 51:1111-1150. Roça do Casal do Meio: GrA-13501 and GrA-13502; Tanchoal: GrA-9572 and GrA-9270; Souto 1: Beta-280041; Paranho, cist 4: GrA-14008; Paranho, cist 4: GrA-14007; Paranho, cist 4: GrA-22445; Paranho, cist 3: GrA22444; Paranho, cist 2: GrA-5412; Paranho, cist 2: GrA-5410; Paranho, cist 1: GrA-5425; Senhora da Ouvida 7: GrA-1251, GrA-1248 and GrA-9741. these chronologies. It is quite possible that the phenomenon is set through a remarkable time span, what also might help to understand the diversity of situations, which do not have thus to be understandable merely by different functions and cultural traditions, nor whatsoever by different economic, social, political and ideological organisation (Cruz, et al., 1998 b: 51). Alongside these small monuments with tumuli, another reality seems to be adopted: small monuments or enclosures have also been built but without tumuli. The best example excavated is Travessa da Lameira de Lobos (Cujó, Castro Daire), flat monument externally defined by a circle of firmly embedded stones (Cruz & Vilaça, 1999: 132; Cruz, 2001: 331). Although with all the differences, namely the type of internal structures, this solution of physically and symbolically demarking the death space by a stone ring was adopted in other occasions such as in Paranho, where a “circular line of stones sealing the enclosure” was to be found (Coelho, 1925: 14). Also with differences that we must acknowledge, slabs placed originally in vertical position and juxtaposed (Cardoso, et al., 1998: 328 e 331) also defined two subcircular structures of Monte de São Domingos (Malpica do Tejo, Castelo Branco). It should also be noted that these stone circles intended to protect the dead might integrate other symbolic markers such as spellers marked in some slabs, as observed in Travessa Lameira de Lobos, with almost all the slabs

decorated either with reticulated compositions or with semi-circles (Santos & Marques, 2007: 39). The presence of decorated slabs in funerary-ritual structures from Late Bronze Age was known from Casinha Derribada 3 (Mundão, Viseu), which central pit with four deposited vases was covered by slab engraved with reticulate (Cruz, et al., 1998b) iconographic matrix that, hypothetically, could be assumed as a metaphor for a web or net symbolically protective of the world of the dead. More certain is the need to carry on the research on this type of monuments which, as seen in these reflections and despite certain cross-cutting features, incorporate multiple situations in the most varied domains. And the truth is that behind an apparent similarity, or an illusory similitude, the solutions appear to have been diverse. Also, it cannot be emphasised enough that, among all those constructive realities, structures exist that are not strictly or exclusively funerary, although they may be related with mortuary rituals and evoke highly complex ceremonials (Vilaça & Cruz, 1999: 87). Despite not having a precise chronology, the interesting case of the small structure of Vale de Mós 1 (Oleiros), recently excavated (Caninas, et al., 2009), featuring a stone pavement surrounded by a peripheral stony ring almost as defining a sort of open platform and which has precisely not shown any evidence of incineration practices fits into

129 

 

these acting scenarios of ritual nature connected with death. On the other hand, after the first discovery and characterisation of several nuclei from the Viseu/Vila Nova de Paiva/Castro Daire region with some sets (or elements of sets) excavated and systematised in due time (Cruz, 2001), many other groups of small tumuli have been found in Aveiro, Coimbra, Guarda and Castelo Branco district (Caninas, et al., 2008; 2009).

The possibility of a fourth origin in the Iberian Peninsula itself, of a multipolar nature, without excluding the others, must not be neglected because it is also necessary to recognise, beyond the unequivocal chronologic discrepancies, that there is not uniformity even in the incineration ritual. Taking the example of a single necropolis, Les Moreres (Crevillante, Alicante), five different forms of incineration dating from the IX-VII centuries BC can be observed, the digger in charge proposing an autochthonous origin to this ritual, from Calcolithic (González Prats, 2002: 391).

The power of fire The importance of fire rituals directly or indirectly associated with funerary practices in Central Portugal Bronze Age materialises in many different ways. The assertion of incineration does not merely translate a ritual change. It is mostly a different paradigm of understanding the materiality of human body which loses importance once it is totally or partially destroyed and not necessarily deposited in its whole or even in part. The focus of attention seems to be transferred from the physical body to its vestigial or immaterial memory, preserved through other practices. It also does not translate abandonment of other practices, like the inhumation rituals present in Roça do Casal do Meio, in Casal de Santo Amaro, or in Medronhal (Condeixa-a-Nova) cave. It is not possible to know if this temporal bi-ritualism (XI/X-VIII centuries BC) would have been practised by culturally distinctive or related communities, nor if any of them prevailed in the following centuries in Central Portugal. In another way, the idea that incineration would be dated from Late Bronze and would go exclusive or fundamentally paired with the “Urnfield” question has to be put in perspective. In European terms, it’s well known that incineration was practised since at least the Early Neolithic (Zammit, 1991: 70, among others), being certain that it would have become common practice mainly, and in certain regions (e.g. Hungry, England), from the Early Bronze Age onwards (Harding, 2003: 120). At peninsular level, the existence of three independent cores in Late Bronze Age – Early Iron Age with respect to the adoption of incineration practices continues to be recognised: one continental, from Catalonia, connected with the “Urnfield” phenomenon; one Mediterranean, Phoenician and/or pre-Phoenician; one Atlantic, arrived by maritime way, as well as the latter (Pellicer Catalán, 2008). The Atlantic origin for the Northwest of Iberia had been previously sustained by Ana Bettencourt, who referred the phenomenon back to the Middle Bronze (Bettencourt, 1995: 113), chronology recently corrected by the researcher herself due to new evidences, pointing towards Late Calcolithic (Bettencourt & Meijide, 2009).

How the question is framed is perhaps impeding it from being fully understood. This is because not all that is new comes from outside, nor everything we know is adopted, nor the problem itself is reduced to the binomial inhumation-incineration. Nor can we approach the problem through radical assumptions of replacing the former by the latter (Vilaça, et al., 1999: 17). Among these two stadiums, there is room for a third or another one which, while not corresponding to formal incinerations, uses and handles fire in inhumation practices. Actually, the material - and symbolic – power of fire acting over the bodies must not be confounded exclusively with body incineration practices, with or without subsequent recollection (total or partial) at urn, pit, small pit, etc. Skeletal remains partially burned or with fire traces have been collected in megalithic monuments as witnessed, for example, by Bola de Cera (Marvão) dolmen (Oliveira, 1998: 448 e 451). These are not mere inhumations; nor incinerations as normally understood. They are both simultaneously, with the fire making the difference, but unable to assert themselves. The question is very much about the fact that fire rituals are not confined to funerary practices directly and exclusively referred to the dead. Beyond them, and with them, fire manipulation must have been much more recurrent than we think, including without protection and fireguard (e.g. spread ashes in land or thrown to water), which makes it often difficult to circumscribe the specific nature of its use. Thus, it is the re-conceptualisation of the concepts of incineration, cremation and fire rituals that must be put on the table. In central Portugal the problem of funerary practices of fire rituals is an open theme. There are no clear pieces of evidence so far to allow its association with megalithic contexts. Some clues point to, as early testimonies, the middle of the II millennium BC, as indicated by the traces identified in Serra da Muna (Viseu) monument 2 (Cruz, et al., 1998 a), but it is possible that cases further back in time may come to be identified. Anyway, more consistent evidences only appear in final Bronze Age contexts. Undoubtedly, the identification and posterior valorisation of Paranho necropolis, as mentioned in the beginning of this paper, is a fundamental reference and a good 130 

 

example of recollection in urn of burned remains, human bones and also artifacts, namely in bronze, as revealed by cist 2’s container (Cruz, 1997: 90). This practice of using a ceramic container to receive human cremation remains, accompanied or not by objects also submitted to fire, is known in other parts of the Portuguese territory, with suggestive distribution along the River Tagus and undoubtedly as a cultural mark of the Late Bronze and Early Iron Age. The “Tagus Line” In the Southern frontier of the region under analysis in this paper is a line gathering some of the most interesting testimonies of incinerations dated from the final Bronze Age or beginning of the next phase. On the whole they are not well known but, their geographical incidence suggests they must have some significance. They cannot obviously be dissociated from the river history as a privileged fluvial-maritime communication hub between the Western Atlantic and the continental inland (Vilaça, 1995: 410-411; Vilaça, et al., 1998: 38; Vilaça & Arruda, 2004: 39) or from the assumed diffusion of incineration rituals through that inland and continental route advocated by some researchers. Among those pieces of evidence are Monte de São Domingos (Malpica do Tejo, Castelo Branco) where two subcircular structures were identified, one of them with

an urn appearing to have contained only human bones (Cardoso, et al., 1998). Walking downstream is the interesting case of mamoa 1 of Souto (Bioucas, Souto, Abrantes) (Cruz, et al., 2011), already recorded. It is an urn placed in a small pit with bone remains from incineration, as well as metal fragments (possible bracelet), perhaps burned with the body, inside of which was a second container also with ashes and human bone remains, plus other organic elements such as seeds. These elements were found in the central area of the small tumulus (6m diameter and about 50cm high) built with pebbles, i.e., the nonmonumental tomb construction tradition has been used. Souto 1, another case of a Late Bronze Age incineration, as revealed by the examination of materials and radiocarbon dating (1125-903 cal BC) (Cruz, 2011: 146), is also one more case of funerary deposition with bone remains collected in urn, although with a spatial structure quite different from those found in Paranho and Monte de São Domingos. Continuing the path through the Middle Tagus, we arrive at Alpiarça, a region where important testimonies can be found which are regularly referred to in the bibliography as “Alpiarça urnfield”, since the first findings dated from 1916 (main references gathered in Vilaça, et al., 1999). It is at least the well-known polynucleated cemeteries of Cabeço da Bruxa, Tanchoal and Meijão whose

Figure 11.4: Urns with calcined bones from Paranho necropolis (Photo by R. Vilaça).

131 

 

importance goes far beyond them, despite the way most of the data have reached the present day. There are still many outstanding issues to be clarified, but it is obvious that this is a completely different case, not only because of the particularity of the constructions and spatial organisation of mortuary deposits, but also because it gathers heterogeneous situations, including at chronological level. The Alpiarça necropolis and burials, revealing some affinities with certain contexts of the socalled “Qurénima group” systematised in due time (Lorrio, 2008), are well worth a joint in-depth reassessment. Concerning only the Portuguese Tagus, beyond the Monte de S. Domingos, Souto 1, Tanchoal, Meijão and Cabeço da Bruxa cases, with objective and minimally safe data, other indicators (unfortunately very badly known) might, as working hypothesis, be associated with the problem under discussion. We refer to presumable funerary depositions related to incineration practices in Quinta da Alorna (Almeirim), where a carinated vase with handle was gathered in unknown circumstances (Schubart, 1971: 166); in Salvaterra, a necropolis has been recorded (Savory, 1951: 375); in Almoster, also a necropolis, from where a complete vase25 of cylindrical neck is known (Savory, 1951: 375; Spindler, et al., 197374: 129); in Santarém (without precise location) referred to as necropolis and burials (Savory, 1951: 375; Spindler, et al., 1973-74: 144).

chronology is not completely answered. As known, the only radiocarbon dates are of one of Tanchoal contexts and put it between middle XI century BC and early IX century BC (Vilaça, et al., 1999). Thus, in its whole, the various necropolises might and should express wider spectrum chronological parameters, since final Bronze to Early Iron Age. The German researchers working in the area had already claimed a higher antiquity to Cabeço da Bruxa necropolis relative to Tanchoal and Meijão, considering the difference between the two legacies (Kalb & Höck, 198182). And the identification, among the collection in “Casa Museu dos Patudos”, of several bowl fragments from “tomb F” (as Gustavo Marques called it26), at Alto do Castelo (Alpiarça) habitat, not only raises new questions at chronological level (including the “Orientalizing” problematic), as well as interesting problems at spatial organisation level, with death “invading” the spaces of the living. These issues as well as the settlement/necropolis binomial problematic (with some interesting data to value) on the Late Bronze Age cannot, however, be developed here due to text size constraints. In short, the whole area of the Portuguese Middle Tagus involving both banks, with all the known data and still to be explored, both in the field of funerary practices and the settlement in general, seems to be one of the strategic regions for the study of social dynamics of the Bronze and Iron Ages within the Portuguese territory and peninsular reach due to the problems involved. On the other hand, and despite the specificities of the reported cases, from the way they were spatially implemented and materialised (with or without reference markers) to the types of structures and depositions of the cremated remains, whether or not associated with materials, there is a common denominator, i.e., the recollection of ceramic containers of the cremated remains (only bones or bones and metallic materials) as a specific practice which consolidated at the turn of the millennium.

Figure 11.5: Urn with ashes and bracelets from Tanchoal necropolis (Photo by A. Roldão, ICBAS). This information does not certainly offer very safe data in most of the cases but the geographical concentration of the finds and its proximity to Alpiarça must have some meaning that deserves further investigation, beginning with its confirmation and proceeding to the chronological assignment. A source of controversy since the moment of the identification of the first testimonies in the 20’s of the last century, the question of Alpiarça funerary world

Although not exclusive from this region, the Tagus Valley seems to have been a privileged region in what concerns the transmission of this new ritual. But, while varying from case to case, the solutions known reflect (and this is the most important) the creative potential of communities. Also here, and albeit within a general common frame defined by the incineration ritual (this one also without tight norms), what stands out are the specificities of each situation. For each case a different approach. The diversity seems to reflect the coexistence of traditions such as small tomb structures like those in Souto 1 alongside with the assimilation of another concept of burial marked by total (?) invisibility and

                                                            

                                                            

It is worth mentioning the particularity of the vase having traces of perforation in the middle of the bulge, maybe of funerary matrix. About the use of perforated vases in funerary contexts see Vilaça and Cruz, 1999: 87, nota 33.

Unique manuscript records by Gustavo Marques are available for consultation in “Casa Museu dos Patudos”. We thank its Director, Dr. Nuno Prates, the permission granted. The materials are being studied by a team coordinated by our colleague Ana Margarida Arruda.

25

26

132 

 

(apparent) absence of structured and independent spaces for each urn, as it seems to have occurred in Alpiarça. The public, unburied and unprotected death At this point, and going back to the beginning, it seems clear that in final Bronze Age death did not disappear from the archaeological record in Portuguese central territory. It is present, materialises in many different ways, and sometimes comes disguised, as illustrated by the above mentioned examples. When it doesn’t come, what inevitably drags us to the field of non-demonstrable archaeology, there is still room for other working hypotheses. In addition to inhumation, there are the incineration and the use of fire rituals with manifestations that are clearly differentiated but not always easy to recover. But perhaps fire has not always been the only responsible for the new forms of “communicating” with death and of instilling in it a marked invisibility. As we have already stressed, death to these societies without writing wouldn’t be an ephemeral fact circumscribed to the purely biological domain, rather a “rite of separation” implying a long and complex process of separation from the body until its transformation, disposal and deposition. Or rather also of exhibition. The practice of exhibiting bodies in scenarios that left no traces is admitted as a very likely hypothesis (Vilaça & Cruz, 1999: 76; Vilaça, 2000: 40). And, contrarily to those rituals with safeguarded inhumations and incinerations, this one has no protection but though it is visible and consequently public. The admissibility of such a subtle ritual, either as a form of final disposal of the bodies (i.e., without burial) or as pre-depositional practice, albeit prolonged, an intermission for the bodyhandling ritual, results from some considerations. From all these considerations mention should be made to the one that suggests the disarticulation of human bones maybe the last link of a prolonged process where death was temporarily exposed and subject to a final fragmentation. This phenomenon, interestingly discussed by Joana Brück in several works (e.g. Brück, 1995: 247; 253, 257), translates not only a new type of practices that are not exactly funerary but involve human remains but also, and

most important, a different approach to death and the human body. It is a fragmented broken death redistributed by different contexts, which gains mobility and starts accompanying the living, instead of them (re)worshiping cyclically in proper spaces. The body is transformed into a good which circulates among people and among places (Fowler, 2004: 40), playing like any other object, an active role in reproduction and social renegotiation. It is known that body exhibition was a ritual, practised by ancient Iberia Peninsula populations, as Vaceus and Celt Iberian, as attested in written (Claudio Eliano X, 22; Silio Itálico Guerra Púnica, III, 340-343) and iconographic sources, where bodies and warriors are devoured by vultures (Alfayé Villa, 2008: 296; Sopeña, 2005: 381). The display of dead bodies, as recently remembered (Esparza et al., 2012: 115), was also practised in other periods and places from the Greek “Obscure Centuries” to Black Africa, from the 1880’s North American prairies to Indian Parsi (Tillier, 2009: 8). The dead were exposed to natural elements to allow their return to nature, on the top of trees or on platforms, conventionally called “platform-tombs” or “air-burials” (Fahlander & Oestigaard, 2008: 6). Therefore, if it is certain that these and those data cannot be directly imported to the peninsular world of 3000 years ago, the truth is that the origin of corpse display rituals may have been much more ancient. In this sense, it seems to attest the strong arguments (e.g. evidence of dog bites) used by the Spanish colleagues in that study to prove that the exposure of dead bodies was the norm among the Cogotas I communities. While unburied and exposed, death acquires, as mentioned before, unequivocal public character propitious for the involvement of the various social players, thus contributing to reinforce the communitarian identity. However, as its display is inversely proportional to its perennity, it would also be brief, setting ephemeral “funerary landscapes”, that we can only suppose. It seems that the more public the death; the more difficult it is to capture the subtlety of its nature. Once the bodies have disappeared, the memory of the deceased is the only thing that subsists. And every dead person should have had a place to inhabit, or to remember… we know, however, that just a few had it.

133 

 

Figure 11.6: Sites referred to in the text (approximate location). The memory (in / of) the places In final Bronze Age there seems to be an unequivocal tendency for a real disidentification of the body and, in particular, the triumph of a different paradigm in understanding the materiality of human body which devalues and loses existence, once it is totally or partially destroyed, namely by fire, being only partly deposited, possibly dismembered, dispersed, unburied. But if the dead body seems to have been physically devalued that was not the case with the dead - they remain in another way. The focus of attention is thus transferred from physical body to the memory of it, which is collectively and socially preserved through other rituals, other material forms, other references. Let us say that, while the physical body is eliminated, the social body is constructed and invented through memory. The social, collective, common, shared memory is, however, a shortterm memory, which is extended by means of different strategies. Among them, and at all times, monument construction stands out and with it the commemorative ceremonies

(Connerton, 1999: 8, 47, 81). The implantation of a hallmark (stele, monolith, statue-menhir, etc.) in a place with meaning, or upon which meaning is conferred by introducing this new scenic element was practised in remote times in the Portuguese Central territory.27 Also in this respect we must distinguish inland areas, where this is observed, from coastal areas, where they are unknown so far and where the strategies would have been different. In those societies without writing, such practice must have been particularly assertive and perennial, because it was engraved in stone - procedure used by the communities to record time and their own history, i.e., knowledge of (and with) its past (Vilaça, 2011: 8). This self-acknowledgement as a social reproducer of memory of sense of place, of belonging, of identification should entail actions - commemorative ceremonies - involving the communities, including the neighbouring ones. The entities that are represented or evoked assume, from this perspective, a collective value and the place where they are inserted become a ‘meeting point’ and a forum for the

                                                             27

Among others, look up more recently Cardoso, 2011; Cruz and Santos, 2011.

134 

 

reproduction of sociability which is cyclically (re)visited (Vilaça, et al., 2011: 310). The communities identify themselves not only with and in the place where they have their dead, but equally with the place where their dead, i.e., their ancestors, could be remembered. Therefore, the place of the dead is also the place where they are evoked socially, communitarianly and publicly. Without signalling the burials, of which there are no direct evidences in the area under analysis (and almost none in the others, but yet still existing), the steles and statue-menhir from the final Bronze Age assume, however, a potentially para-funerary character, in the sense that power and death, ancestry and memory may be confounded, or meet together. Standing in passage or border areas, or in meeting places for neighbours, sometimes under the towering gaze of the village which is not far,28 the final Bronze Age steles, without being specifically funeral, cannot be excluded from the death discussion problem. Final notes: the “faces of death” From the elements described (or remembered) and the comments developed in this text, we may conclude that as certain as death is the archaeological evidence of its presence in central Portugal in final Bronze Age, as demonstrated by the various radiocarbon datings (Chart 12.1), despite the generic trend to a certain “dematerialisation” of the human body. And it is equally certain that, in this region, it was expressed in a variable, irregular form without a general rule. The bi-ritualism – inhumation and incineration – is one of the most expressive elements in this domain, but many others were herein underlined. Such is the case with the indelible presence of small thousand-year tumuli shaping true “patrimonialised” landscapes. We have focussed on data, questions, working hypotheses, diverse practices and identitarian discourses, some more linked to the past, others assuming real breakthroughs in terms of space, contexts and rituals. Nor have we dwelled on the detail that would have merited the question of the so-called “grave goods”, which is little exuberant, scarce and even absent from some known contexts. Death seems to be less “objectified”. Or has it repeatedly been so in deposits, especially metallic? When present, most of the assets are ceramics and some metal that, at the time, was marginal in explicit funerary context.

adornments predominate, which reflects the importance of the individual as a person. The presence of weapons is negligible, which contrasts with the preference for personal adornment and ornamentation, especially bracelets (some of them also submitted to fire), revealing well that certain materials would have carried higher semiotic weight than others. From the social point of view, the trend towards an individual treatment of death with an increasing personal mark is unequivocal. Even so, cemeteries with different burials prevail, perhaps of relatives. The parental relations seem to be thus dominant. But we cannot discard other types of relationships such as bonds of loyalty, which would come to mark the following periods. Investigating the problematic of the so-called “dead for accompaniment” (Testart, 2009), either voluntary or imposed by social precepts, is open field for debate as is the case with the two men burial from Roça do Casal do Meio. In summary, the communities have given diverse responses to the challenge of death. It is not easy to find the motives for such different behaviours and attitudes towards death. Environmental constraints and different economies, traditions, beliefs and influences, social, age and gender precepts, “marital status” – a married woman distinguishes herself from another who is single and a woman who became a mother differs from all the others – types of death (accidental, natural, inflicted, in action); times of change, when the “Other” was firmly affirmed including by its presence, by increasing travels and contacts: here are some hypotheses to be explored, case by case, which determined what to do with and how to dispose of the human body, materially inert but with an enormous symbolic power. Therefore, it is crucial to go on researching because data and contexts will never be sufficient to get closer to understanding how the living dealt with their dead, conceived the world and created their own mindset. It is also in this sense that we do not subscribe an absolute “archaeology of the dead”, but of the living (even if death is the centre of concern) focusing on their actions, options, illusions and social strategies. In its worldview, death is ubiquitous. And, in central Portugal, it expresses in a wide variety of ways.

As known, this was instead left apart in other types of context, the so-called “deposits”. Yet still, the metal is present in funerary burials as diverse as Paranho, Roça do Casal do Meio, Alpiarça, Medronhal, Souto 1, etc. And it does not exclude gold, as suggested by the Casal de Santo Amaro necklace. In all of them metallic personal

                                                             28

Case of Telhado stele (Fundão), study in progress by the author, João Mendes Rosa and Joana Bizarro.

135 

 

References ALFAYÉ VILLA, S. (2008) – Iconografia, identidade y sociedade. Gallaecia. Santiago de Compostela: Universidade de Santiago de Compostela, nº 27, p. 285-304. BELÉN, M.; ESCACENA, J.L.; BOZZINO, M.I. (1991) – El mundo funerario del Bronce Final en la fachada atlantica de la Peninsula Iberica. I. Analisis de la documentation. Trabajos de Prehistoria. Madrid: Consejo Superior de Investigaciones Científicas, volume 48, p. 225-256. BETTENCOURT, A. (1995) – Dos Inícios aos Finais da Idade do Bronze no Norte de Portugal. A Idade do Bronze em Portugal. Discursos de poder. Lisboa: SEC/IPM/MNA, p. 110-115. BETTENCOURT, A.; MEIJIDE CAMESELLE, G. (2009) – Agro de Nogueira, Melide, A Coruña: novos dados e novas problemáticas. Gallaecia. Santiago de Compostela: Universidade de Santiago de Compostela, nº 28, p. 33-40. BRADLEY, R. (1998) – The Significance of Monuments. On the shaping oh human experience in Neolithic and Bronze Age Europe. London: Routledge. BRÜCK, J. (1995) – A place for the dead: the role of human remains in Late Bronze Age Britain. Proceedings of the Prehistoric Society. Cambridge: Cambridge University Press, volume 61, p. 245-277. CALDEIRA, D. (2012) – Arqueologia de Alpiarça: o caso dos chamados “Campos de Urnas”. Revisão, problemáticas e perspectivas. Coimbra: Universidade de Coimbra, Master dissertation. CANINAS, J.C.; HENRIQUES, F.; BATISTA, A.; MONTEIRO, M.; CHAMBINO, M.; HENRIQUES, F.R.; CANHA, A.; CARVALHO, L. (2009) – Estruturas monticulares antigas na fronteira Sul do concelho do Sabugal. Sabucale. Sabugal: Museu Municipal, nº 1, p. 21-38. CANINAS, J.C.; SABROSA, A.; HENRIQUES, F.; MONTEIRO, J.L.; CARVALHO, E.; BATISTA, A.; F.R.; CHAMBINO, M.; HENRIQUES, MONTEIRO, M.; CANHA, A.; CARVALHO, L.; GERMANO, A. (2008) – Tombs and rock carvings in the Serra Vermelha and Serra de Alvéolos (Oleiros, Castelo Branco). In BUENO-RAMIREZ, P.; BARROSO-BERMEJO, R.; BALBÍN BERHMANN, R. (eds.) – Graphical Markers and Megalith Builders in the International Tagus, Iberian Peninsula. Oxford: BAR Publishing, BAR International Series 1765, p. 89-102. CARDOSO, J.L. (1999-2000) – Aspectos do povoamento da Baixa Estremadura no decurso da Idade do Bronze. Estudos Arqueológicos de Oeiras. Oeiras: Câmara Municipal, nº 8, p. 355-414. CARDOSO, J.L. (2000a) – Manifestações funerárias da Baixa Estremadura no decurso da Idade do Bronze e da Idade do Ferro (II e I milénios A.C.): breve síntese. Actas do 3.º Congresso de Arqueologia Peninsular. Porto: Adecap, volume V, p. 61-79.

CARDOSO, J.L. (2000b) – A sepultura da Roça do Casal do Meio (Sesimbra) no quadro dos rituais funerários da Idade do Bronze da Baixa Estremadura. Discursos. Língua, Cultura e Sociedade. Lisboa: Universidade Aberta, III série, nº 2, p. 243-251. CARDOSO, J.L. (2011) – A estela antropomórfica de Monte dos Zebros (Idanha-a-Nova): seu enquadramento nas estelas peninsulares com diademas e “colares”. In VILAÇA, R. (coord.) – Estelas e estátuas-menires. Da Pré à Proto-história. Sabugal: Câmara Municipal do Sabugal/Centro de estudos, Actas das IV Jornadas Raianas, p. 89-116. CARDOSO, J.L.; CANINAS, J.C.; HENRIQUES, F. (1998) – Duas cabanas circulares da Idade do Bronze Final do Monte de São Domingos (Castelo Branco). Estudos Pré-históricos. Viseu: Centro de Estudos da Beira Alta, volume VI, p. 325-345. COELHO, J. (1925) – A Necrópole do Paranho, Viseu. Tipografia popular: author. CONNERTON, P. (1999) – Como as Sociedades Recordam. Oeiras: Celta editora. CRUZ, A. R. (2011) – A Pré-história Recente no Vale do Baixo Zêzere, Arkeos, 30, CEIPHAR, Tomar. CRUZ, A. (2011) – A Pré-História Recente do vale do baixo Zêzere. Tomar: Centro Europeu de Investigação da Pré-História do Alto Ribatejo, Arkeos, perspectivas em diálogo, volume 30, dissertação para obtenção do grau de Doutor, Universidade de Trás-os-Montes e Alto Douro, 273 p. CRUZ, A. R.; GRAÇA, A.; BATISTA, A. (2011) – Recente Prehistory and Protohistory in Abrantes and Constância council (Portuguese Middle Tagus) – The research preliminary state. In BUENO, P.; CERRILLO-CUENCA, E.; GONZÁLEZ CORDERO, A. (eds.) – From the Origins: the Prehistory of Inner Tagus Region. Oxford: BAR Publishing, BAR International Series 2219, p. 93-109. CRUZ, D. J. (1997) – A necrópole do Bronze Final do “Paranho” (Molelos, Tondela, Viseu). Estudos Préhistóricos. Viseu: Centro de Estudos da Beira Alta, volume V, p. 85-109. CRUZ, D. J. (1998) – Expressões funerárias e cultuais no Norte da Beira Alta (V-II milénios a.C.). Estudos Pré-históricos. Viseu: Centro de Estudos da Beira Alta, volume VI, p. 149-166. CRUZ, D. J. (1999) – A necrópole do Bronze Final do “Paranho” (Molelos, Tondela). Resultados das datações radiocarbónicas. Estudos Pré-históricos. Viseu: Centro de Estudos da Beira Alta, volume VII, p. 263-270. CRUZ, D. J. (2001) – O Alto Paiva: megalitismo,

136 

diversidade tumular e práticas rituais durante a PréHistória Recente. Coimbra: Faculdade de Letras da Universidade de Coimbra, 2 volumes, PhD dissertation. CRUZ, D. J.; VILAÇA, R. (1999) – O grupo de tumuli da Senhora da Ouvida (Monteiras/Moura Morta, Castro Daire, Viseu). Resultados dos trabalhos arqueológicos. Estudos Pré-históricos. Viseu: Centro de Estudos da Beira Alta, volume VII, p. 129-161. CRUZ, D. J.; SANTOS, A. (2011) – As estátuas-menires da serra da Nave (Moimenta da Beira, Viseu) no contexto da ocupação pré-histórica do Alto Paiva e da Beira Alta. In VILAÇA, R. (coord.) – Estelas e estátuas-menires. Da Pré à Proto-história. Sabugal: Câmara Municipal do Sabugal/Centro de estudos, Actas das IV Jornadas Raianas, p. 117-142. CRUZ, D. J.; GOMES, L.F.; CARVALHO, P.S. (1998a) – Monumento 2 da Serra da Muna (Campo, Viseu). Resultados preliminares dos trabalhos de escavação. Estudos Pré-históricos. Viseu: Centro de Estudos da Beira Alta, volume VI, p. 375-395. CRUZ, D. J.; GOMES, L. F.; CARVALHO, P. S. (1998b) – O grupo de tumuli da Casinha Derribada (concelho de Viseu). Resultados preliminares da escavação arqueológica dos monumentos 3, 4 e 5. Conimbriga. Coimbra: Instituto de Arqueologia da Faculdade de Letras de Coimbra, nº 37, p. 5-76. CRUZ, D. J.; VILAÇA, R.; SANTOS, A. T.; MARQUES, J. N. (2000) – O grupo de tumuli do Pousadão (Vila Nova de Paiva, Viseu). Estudos Préhistóricos. Viseu: Centro de Estudos da Beira Alta, volume VIII, p. 125-150. CUNHA, E.; FERREIRA, M. T., WASTERLAIN, S. (2008) – Intervenção de Antropologia Biológica. Gruta do Algarinho. Coimbra: Relatório de trabalhos antropológicos. ESPARZA ARROYO, A.; VELASCO VÁZQUES, J.; DELIBES DE CASTRO, G. (2012) – Exposición de cadáveres en el yacimiento de Tordillos (Aldeaseca de la Frontera, Salamanca). Perspectiva bioarqueológica y posibles implicaciones para el estúdio del ritual funerário de Cogotas I. Zephyrus. Salamanca: Universidad de Salamanca, volume LXIX, p. 95-128. FAHLANDER, F.; OESTIGAARD, T. (2008) – The materiality of death: bodies, burials and beliefs. In FAHLANDER, F.; OESTIGAARD, T. (ed.) – The Materiality of Death. Oxford: BAR Publishing, BAR International Series 1768, p. 1-16. FIGUEIRAL, I. (1997) – Necrópole do Paranho (Molelos, Tondela). Resultados da análise dos carvões vegetais. Estudos Pré-históricos. Viseu: Centro de Estudos da Beira Alta, volume V, p. 121122.

FIGUEIREDO, E.; ARAÚJO, F.; SILVA, R. (2011) – A ponta de lança da Gruta da Nascente do Algarinho (Penela) no conjunto da metalurgia do Bronze Final. Actas do Encontro Internacional sobre Ciências e Novas tecnologias Aplicadas à Arqueologia na villa romana do Rabaçal, Penela, Terras de Sicó, Portugal. Penela: Câmara Municipal de Penela, p. 41-49. FONTIJN, D. (2007) – The significance of ‘invisible’ places. World Archaeology. London: Routledge, volume 39, issue 1, p. 70-83. FOWLER, C. (2004) – The Archaeology of Personhood. London: Routledge. GARCÍA SANJUÁN, L. (2005) – Las piedras de la memoria. La permanencia del megalitismo del Suroeste de la Península Ibérica en el II y I milénios ANE. Trabajos de Prehistoria. Madrid: Consejo Superior de Investigaciones Científicas, volume 62, nº 2, p. 85-119. GONZÁLEZ PRATS, A. (2002) – La necropolis de cremación de Les Moreres (Crevillente, Alicante, Eapaña – s. IX-VII AC). Alicante: III Seminario Internacional sobre Temas Fenícios. HARDING, A. F. (2003) – Sociedades europeas en la Edad del Bronce. Barcelona: Ariel Prehistoria. KALB, Ph. (1994) – Reflexões sobre a reutilização de necrópoles megalíticas na Idade do Bronze. Estudos Pré-históricos. Viseu: Centro de Estudos da Beira Alta, volume II, p. 415-426. KALB, Ph.; HÖCK, M. (1979) – Escavações na necrópole de mamoas “Fonte da Malga” – Viseu, Portugal. Beira Alta. Viseu: Assembleia Distrital de Viseu, volume XXXVIII, nº 3, p. 593-604. KALB, Ph.; HÖCK, M. (1981-82) – Cabeço da Bruxa, Alpiarça (Distrito de Santarém). Relatório preliminar da escavação de Janeiro e Fevereiro de 1979. Portugália. Porto: Universidade do Porto. Faculdade de Letras, Nova Série, volume 2-3, p. 61-69. LORRIO, A. J. (2008) – Qurénima. El Bronce Final del Sureste de la Península Ibérica. Madrid: Real Academia de la Historia/Universidade de Alicante. LUCAS, G. (2005) – The Archaeology of Time. London: Routledge. PELLICER CATALÁN, M. (2008) – Los inícios del rito funerário de la incineración en la Península Ibérica, Tabona. Sevilla: Universidad de Sevilla, nº 16, p. 1335. PEÑALVER, X. (2005) – Los crómlech pirenaicos. Bolskan. Huesca: Instituto de Estudios Altoaragoneses, nº 22. SANTOS, A.T.; MARQUES, J.N. (2007) – Os tumuli do Rochão (Castro Daire, Viseu). Conimbriga. Coimbra: Instituto de Arqueologia da Faculdade de Letras de Coimbra, nº 46, p. 27-51. 137 

SAVORY, H.N. (1951) – A Idade do Bronze Atlântico no Sudoeste da Europa. Revista de Guimarães. Guimarães: Sociedade Martins Sarmento, volume 61, nº 3-4, p. 323-377. SCHUBART, H. (1971) – Acerca de la cerámica del Bronce Tardio en el Sur y Oeste Peninsular. Trabajos de Prehistoria. Madrid: Consejo Superior de Investigaciones Científicas, volume 28, p. 153182. SILVA, A. M.; CUNHA, E. (1997) – As incinerações da Necrópole do Paranho: abordagem antropológica. Estudos Pré-históricos. Viseu: Centro de Estudos da Beira Alta, volume V, p. 111-119. SOPEÑA, G. (2005) – Celtiberian ideologies and Religion. E-Keltoi. Journal of Interdisciplinary Celtic Studies. Milwaukee: University of Wisconsin – Milwaukee Center for Celtic Studies, nº 6, p. 347388. SPINDLER, A.; BRANCO, A.C.; ZBYSZEWSKY, G.; FERREIRA, O.V. (1973-74) – Le monument à coupole de l’âge du Bronze final de la Roça do Casal do Meio (Calhariz). Comunicações dos Serviços Geológicos de Portugal. Lisboa: Serviços Geológicos de Portugal, nº LVII, p. 91-154. TESTART, A. (2009) – Partir dans l’au-delà accompagné ou le rôle des fidélités personnelles dans la genèse du pouvoir. In GUILAINE, J. (dir.) – Sépultures et sociétés. Du Néolithique à l’Histoire. Paris: Éditions Errance, Séminaire du Collège de France, p. 71-80. TILLIER, A.-M. (2009) – L’homme et la mort. L’émergence du geste funéraire durant la préhistoire. Paris: CNRS Éditions. VILAÇA, R. (1995) – Aspectos do povoamento da Beira Interior (Centro e sul) nos finais da Idade do Bronze. Lisboa: IPPAR, Trabalhos de Arqueologia, 9, 2 volumes. VILAÇA, R. (1999) – Some comments on the archaeological heritage of the Late Bronze Age in Beira Interior. Journal of Iberian Archaeology. Porto: Universidade do Porto, nº 1, p. 173-184. VILAÇA, R. (2000) – Notas soltas sobre o património arqueológico do Bronze Final da Beira Interior. In FERREIRA, M.C., PERESTRELO, M.S., OSÓRIO, M., MARQUES, A. (eds.) – Beira Interior. História e Património. Guarda: Actas das I Jornadas de Património da Beira Interior, p. 31-50. VILAÇA, R. (2008) – No rasto do Bronze final do

Centro-sul da Beira Litoral: artefactos metálicos e seus contextos. In CALLAPEZ, P.M.; ROCHA, R.; MARQUES, J.; CUNHA, L.; DINIS, P. (eds.) – A Terra: conflitos e ordem. Homenagem ao Professor Ferreira Soares. Coimbra: Universidade de Coimbra, p. 75-88. VILAÇA, R. (coord.) – Estelas e estátuas-menires. Da Pré à Proto-história. Sabugal: Câmara Municipal do Sabugal/Centro de estudos, Actas das IV Jornadas Raianas. VILAÇA, R. (in press) – Da morte e seus rituais em finais da Idade do Bronze no Centro de Portugal: 20 anos de investigação. Estudos Pré-históricos. Viseu: Centro de Estudos da Beira Alta, nº 17, 2012. VILAÇA, R. E ARRUDA, A.M. (2004) – Ao longo do Tejo, do Bronze ao Ferro. Conimbriga. Coimbra: Instituto de Arqueologia da Faculdade de Letras de Coimbra, nº XLIII, p. 11-45. VILAÇA, R.; CRUZ, D. J. (1999) – Práticas funerárias e cultuais dos finais da Idade do Bronze na Beira Alta. Arqueologia. Porto: GEAP, nº 24, p. 73-99. VILAÇA, R.; CUNHA, E. (2005) – A Roça do Casal do Meio (Calhariz, Sesimbra): novos contributos. Almadan. Almada: Centro de Arqueologia de Almada, II série, nº 13, p. 48-57. VILAÇA, R.; CRUZ, D. J.; GONÇALVES, A. H. B. (1999) – A necrópole de Tanchoal dos Patudos (Alpiarça, Santarém). Conimbriga. Coimbra: Instituto de Arqueologia da Faculdade de Letras de Coimbra, nº XXXVIII, p. 5-29. VILAÇA, R.; SANTOS, A.; PORFÍRIO, E.; MARQUES, J.; CANAS, N. (1998) – Lugares e caminhos no mundo pré-romano da Beira interior. Cadernos de Geografia. Coimbra: Departamento de Geografia da Faculdade de Letras da Universidade de Coimbra, nº 17, p. 35-42. VILAÇA, R.; SANTOS, A.; GOMES, S. (2011) – As estelas de Pedra da Atalaia (Celorico da Beira, Guarda) no seu contexto geo-arqueológico. In VILAÇA, R. (coord.) – Estelas e estátuas-menires. Da Pré à Proto-história. Sabugal: Câmara Municipal do Sabugal/Centro de estudos, Actas das IV Jornadas Raianas, p. 293-318. ZAMMIT, J. (1991) – Les sépultures préhistoriques et le feu: utilisation rituelle, crémations et incinérations. Bulletin de la Société Préhistorique Farnçaise. Nanterre: Société Préhistorique Française, tome 88, nº 3, p. 70-72.

138