understanding early metal in Western Europe

Transcription

understanding early metal in Western Europe
Metallurgical Networks and Technological Choice:
understanding early metal in Western Europe
Ben Roberts
Department of Prehistory and Europe
British Museum
Great Russell Street
London WC1B 3DG
[email protected]
Introduction
This paper analyses how metal objects and metal production practices in Western
Europe during the fourth and third millennia BC were transmitted in space and time
and explores how prehistoric communities shaped this process. This encompasses the
earliest copper, arsenical copper, gold, silver and lead through to the earliest coppertin objects in Spain, France, Holland, Belgium, Britain and Ireland. The ideas,
techniques and projects that shaped past scholarship are assessed in order to explore
an approach that goes beyond the limitations of current perspectives. The paper then
evaluates the earliest dates for metal objects and production practices within the
broader European context as well as how the transmission of metal could have
occurred by exploring the presence of earlier comparable or transferable knowledge,
expertise or material practices. The broader spatial and temporal patterning in the
available choices and identifiable actions that influence the production, circulation
and deposition of metal objects in Western Europe are then examined. It can be
shown that, despite the necessary existence of metallurgical networks and centres
enabling production, metal objects very much reflected the specific and changing
demands of communities involved.
1
Chronologies, Migrations and Technologies
Debates concerning the potential antiquity of metal in Western Europe crystallised
during the mid-late 19th century due to the excavation of prehistoric sites containing
copper objects, most famously at Los Millares, southeast Spain (Siret & Siret 1887) as
well as stone tools associated with “primitive mines” in copper ore deposits
(Domergue 1987). The perceived importance of the earliest metals led to the proposal
of a ‘Copper’, ‘Eneolithic’ or ‘Chalcolithic’ Age (e.g. Much 1885; Cartailhac 1886;
Lichardus & Echt 1991) distinct from Ages of Stone and Bronze that achieved
gradual acceptance throughout much of continental Europe though its existence is still
disputed in Britain, Ireland, the Low Countries and Scandinavia. The appearance of
copper objects in the archaeological record therefore heralded a perceived significant
transition from the Stone Age to the Metal Ages. This technological and
chronological shift encouraged broader societal changes in the minds of many
scholars. Consequently, the explanations proposed for the appearance of metal objects
and mines invariably involved technologically superior colonisers – migrating peoples
from the east who settled in western lands bringing their culture, which included
knowledge of metallurgy, with them. Metal was understood throughout all antiquarian
scholarship to have been inherently desirable due to being a self evidently more
advanced and superior technology.
The significant increase in the quantity of archaeological excavations and material
culture studies during the early to mid 20th century had little influence on these
migration-orientated interpretations. Theories therefore remained firmly orientated
towards the influx of metallurgically-skilled outsiders who exploited the local ore
sources. These ranged from the migration or invasion of ‘peoples’ to the more
2
piecemeal and gradual diffusion by small groups and individuals (see Roberts in press
a). Those involved in the spreading and development of the new technology were
frequently seen as having had special status. V. Gordon Childe famously articulated
this in his highly influential concept of itinerant metal smiths (e.g. Childe 1930) that
lay at the heart of his schemes of social complexity and class construction in later
European prehistory (see Rowlands 1971; Wailes 1996).
The application of radiocarbon dating in the mid 20th century enabled the first
independent chronology for early metal objects and metal production. Colin Renfrew
used these to challenge the established ideas of ex oriente lux (e.g. Renfrew 1967,
1973) arguing instead for the independent discovery of metallurgy in southern Iberia
on the basis that the radiocarbon dates were apparently so much earlier than for its
neighbouring regions that the only explanation must be indigenous development. In
other regions of Western Europe, the discussion of origins has concentrated upon
neighbouring regions where there is earlier evidence of metal objects or metal
production with diffusionist models for the earliest metal use remaining uncontested.
The mid 20th century also saw the earliest systematic archaeometallurgical
investigations in Western Europe. These were motivated by the desire to provenance
prehistoric copper, bronze, and latterly gold objects to a particular region, or more
specifically an ore source. Programmes of compositional analysis were established
such as the vast Stuttgart based Studien zu den Anfängen der Metallurgie (S.A.M.)
project for copper, copper alloys and gold (Junghans et al. 1960, 1968, 1974;
Hartmann 1970; 1979; 1982). Unfortunately, the inability to match many metal object
compositions to ore sources reliably meant that the provenance projects were unable
3
to fulfil their original purpose (see Tylecote 1970). Nonetheless, researchers sought to
use the data to address whether there were distinctive metal composition groupings in
space and time (cf. Butler & Van der Waals 1964; Waterbolk & Butler 1965) and
whether these could be equated with the defined cultural groupings such as the Bell
Beaker culture which stretched throughout central and western Europe during the later
third millennium BC (e.g. Case 1966; Harrison 1974). During the 1970s with the
integration of archaeological fieldwork and archaeometallurgical analysis on projects
that were specifically directed towards the investigation, recording and dating of
features related to metal production in a locale or region. The purpose now was to
reconstruct the production of copper objects in order to investigate the raw materials,
technology and techniques that were used. The earliest and most influential of these
projects was an ambitious survey of the mining and metallurgy evidence in the
Huelva area of southwest Spain (Rothenberg & Blanco-Freijeiro 1980; 1981). The
data that this and subsequent projects have generated inspired experimental
replications that could be informed by and compared to archaeological and
archaeometallurgical data (e.g. Happ et al. 1994; O’Brien 2004; Rovira & Guttierez
2005; Timberlake 2005; 2007). These address fundamental technological questions
surrounding metal objects: what they were made of, how they were made, and where
the ore came from (e.g. Tylecote 1987; Craddock 1995).
Whilst new methods of archaeological excavation, archaeometallurgical analysis and
radiocarbon dating all demonstrably transformed the investigation and interpretation
of early metal objects and metal production, the case for archaeological theory beyond
culture-historical and technological/functional perspectives is far less convincing. As
many prehistorians in Western Europe appear unable to dismiss entirely the validity
4
of ‘cultures’, refuse to re-define it or introduce an alternative, no new perspectives are
introduced. Though social evolutionary models are occasionally promised, the
outcome tends to consist of detailed descriptions of technological changes through
time that are argued to signify increasing social complexity (cf. Childe 1944; e.g.
Strahm 1994). More subtle perspectives have argued that metals were prestigious and
tradable as well as being symbols of elite power and potentially subject to elite
control, but they were not produced by full time specialists and did not play a crucial
role in social change or the growth of social complexity in prehistoric societies (see
Gilman 1996). The strong trend towards researching and modelling at local and
regional dynamics of early metal (e.g. O’Brien 2004) has meant that broader
perspectives have been largely descriptive rather than analytical. They compare the
exploitation of different ore sources and the presence of specific production
techniques and objects forms. It is argued that this approach provides few
explanations regarding the presence or nature of metal and is unable to explore how
prehistoric communities shaped this process.
To understand the transmission of metal objects and metal production practices
through space and time requires going beyond looking at the physical and chemical
conditions of production or the final properties of the object to examine and compare
the available choices and subsequent decisions that influenced the creation, use and
deposition of metal. The knowledge, skills and tools that would be required to
perform each identifiable transformation need to be assessed to analyse how variation
and change could have been enacted. In particular, it is important to ask whether
existing circumstances and technologies could have led to independent discoveries of
metallurgy, or if sufficient metallurgical inspirations could have derived from contact
5
with a metal object or knowledge of potential ores and the involvement of fire, or if
the process required gaining expertise from experienced individuals. In order to
explore the nature of early metal, a systemic approach that encompasses each aspect
of early metal creation, use and deposition to reveal a biography (cf Gosden &
Marshall 1999) that is general rather than specific is required (cf. Lechtman 1977;
1996; Hosler 1995; Ottaway 1994, 2001; Killick 2001; Fontijn 2002/3; Needham
2004; Ehrhardt 2005). This would entail the selection of an ore or ore source; the ore
extraction, processing and distribution; the smelting, melting and alloying; the
casting, manipulation and design of objects; the potential object uses; the circulation
of the objects; the extent of object recycling/re-melting; and the nature of object
deposition. Rather than being connected through linear sequences, this approach
emphasises the dynamic nature of the many different inter-relationships existing
between the different aspects (cf. Kingery 1993; 1996; 2001; Knappet 2005). The
actions taken and not taken, the knowledge, skills and tools needed to perform the
transformation and the consequences of decisions can be assessed. The purpose is to
obtain insights into the broader spatial and temporal patterning in the choices and
actions underpinning the changes and continuities in transmission reflected in early
metal and analysing these in the context of the socio-economic and ideological
dynamics that underlay these prehistoric societies.
When is the earliest metal in Western Europe?
To analyse the appearance of early metal in Western Europe, it is necessary to address
metallurgical origins from the broader European perspective and assess the earliest
evidence for gold, lead, silver, copper-arsenic, tin-bronze as well as copper (cf.
6
Kassianidou and Knapp 2005; Ottaway and Roberts 2008). It is during the mid 6th
millennium BC that the earliest native copper and copper oxide exploitation is found
in southeast Europe on sites such as Lepenski Vir (Srejovic 1972) and Divostin in
Serbia (Glumac 1988) and Durankulak in Romania (Todorova 1999). The dating of
copper ore extraction at Rudna Glava, Serbia to the late 6th millennium BC (Jovanovic
1991; Pernicka et al. 1993; Boric and Jovanovic in prep) and at Ai Bunar, Bulgaria, to
the early 5th millennium (Chernykh 1978) together with many contemporary native
copper or copper oxide beads, hooks, needles and awls at sites as far west as
Neszmely, Hungary (Bognar-Kutzian 1976) and Coka, Serbia (Bailey 2000) confirms
relatively extensive exploitation during this period (see Thornton 2002; Krause 2003;
Zachos 2007 for reviews).
The earliest smelting of copper ores in is less clear due to the ephemeral nature of the
evidence, the relative lack of analyses to determine whether smelting took place, and
the inability to distinguish smelted copper from native copper (see Wayman & Duke
1999). There is analysed copper smelting slag at Belovode, Serbia dating to the late
6th millennium BC (Radivojevic 2007) and in the early 5th millennium BC at Selevac,
Serbia (Glumac & Todd 1991) and Promachon-Topolnica (Koukouli-Chrysanthaki &
Bassiakos 2002) on the border of Bulgaria and Greece together with unanalysed
material interpreted as smelting slag throughout southeast Europe at sites dating to the
early 6th- 5th millennium BC (see Glumac & Todd 1991; Thornton 2002; Zachos 2007
for a review). Further west, the excavation of a cemetery at Zengövákony in
southwest Hungary revealed small pieces of smelting slag in a grave. Though
excavated in the pre-radiocarbon era, the subsequent radiocarbon dating of the
ceramic sequence places it in the late 5th millennium BC (Glumac & Todd 1991, 14).
7
The earliest exploitation and working of gold occurs in southeast Europe and
subsequently spreads throughout eastern Europe during the mid 5th – 4th millennium
BC, as demonstrated most spectacularly in the objects adorning the burials at Varna in
eastern Bulgaria (Renfrew 1986; Makkay 1991). In contrast, the earliest gold objects
found to the east and west only date to the 4th and early 3rd millennium BC (Gopher et
al. 1990; Primas 1995). Similarly, the earliest appearance of silver objects occurs in a
spectacular deposit from Alepotrypa cave in southern Greece also dated to the mid 5th
– early 4th millennium BC (Muhly 2002), pre-dating the silver objects found in
Bulgaria and Rumania from the 4th millennium BC. The earliest evidence for tin
bronze is a single piece of slag from Hungary, reportedly dating to the late 5th
millennium BC (Glumac & Todd 1991). However, as this find predates the earliest
occurrences of copper-tin alloys in this region by over a millennium, it may well be an
isolated case of accidental mixed smelting of copper ores with other tin-bearing ores.
In central Europe, the recent excavations at Brixlegg, Austria revealed a small hearth
where copper fahlores that had been subjected to heat treatment, a small quantity of
smelting slag and a copper strip and bead. The radiocarbon dates for the layer (SE 6)
gave a broad range of 4500-3650 cal BC (Bartleheim et al. 2002; Höppner et al.
2005). However, the ceramic material recovered is typical of the metal importing
Münchshofen culture dated c. 4500-4000 cal BC (Matuschik 1992; Nadler & Zeeb
1994), which makes a date of the later 5th millennium BC more probable. This makes
Brixlegg contemporary with, rather than later than, the earliest evidence for sulphidic
ore smelting further east as demonstrated by the recent slag analyses in east Bulgaria
on samples dated to the late 5th to early 4th millennium BC (Ryndina et al. 1999). It is
also far earlier than the subsequent smelting sites as at Götchenberg, Austria dating to
8
3580-3370 cal BC (Lippert 1992; see Krause 2003 for review). In contrast, the
evidence of copper objects in the Pfyn (c. 3800-3300 BC) and Cortaillod (c. 40003300 BC) cultures of the western Alpine region also dates significantly later than
Brixlegg (Strahm 1994; 2005). Debates concerning the extent of primary as opposed
to secondary metal production during this period remain unresolved after recent
critical re-evaluations (see Fasnacht 1991; 1995; Rehren in press).
To the south of the Alpine region, there is tentative but growing evidence to suggest
that copper objects produced through smelting and dating to the 5th millennium BC
were present in northern Italy in addition to the three copper flat axes originally
proposed by Barfield (1966, 63). Though doubts have been expressed regarding these
due to the lack of contextual information (Skeates 1994, 9-12), the recently
discovered copper objects in 5th millennium BC (Mid-Late Neolithic) contexts at sites
in northeast Italy such as Bannia-Palazinne di Sopra and S. Andrea di Travo
(Giumlia-Mair 2005) and the dating of the awl found at the Arene Candide site in
northwest Italy to late 5th millennium/earliest 4th millennium BC (Skeates 1994;
Maggi & Pearce 2005) demonstrate more convincingly the presence of metal during
this period. The recent radiocarbon dating of mining tools found at Monte Loreto in
northwest Italy places copper ore extraction during the mid 4th millennium cal BC
(Maggi & Pearce 2005). Further west in Sardinia, copper and silver objects and
evidence for copper and silver smelting at sites such as Su Coddu are known from the
Ozieri and sub-Ozieri cultural phases which span the late 5th to the later 4th
millennium BC at sites such as Grotta sa Korona di Monte Majore and Cuccuru
Arrius (Lo Schiavo 1988; Lo Schiavo et al. 2005), whilst in Corsica at the site of
Terrina IV, copper objects and copper smelting has been radiocarbon dated to the later
9
4th millennium cal BC (Camps et al. 1988). However, barring two objects in northern
Italy, no silver objects have been found beyond Sardinia in the central Mediterranean
or Western Europe until the end of the 3rd millennium BC (Primas 1995).
In southeast France, the copper awls, dagger and awl fragments and lead beads were
found in contexts at the site of Roquemengarde (Guilaine 1991) that have been
radiocarbon dated to the later 4th millennium cal BC. However, the occasional
discoveries of objects more typical of other regions with older metallurgical traditions
to the east, such as northern Italian ‘Remedello’ style daggers, western Swiss style
spiral pendants and even a gold repoussé diadem whose closest parallels are in the
Balkans, raises the possibility of earlier dates (Elèure 1982, 56; Guilaine & Eluère
1997, 176; Ambert & Carozza 1998, 160-1). This pre-dates the earliest evidence for
copper ore extraction at Vallarade, Cabrières from c. 3100 cal BC which is virtually
contemporary with the earliest copper smelting at nearby La Capitelle du Broum
(Ambert et al. 2002; 2005; Bouquet et al. 2006) and Pioch Farrus 448 (Espérou et al.
1994). The analysis of the slag, slagged ceramics and partially reduced ores have
revealed the significant presence of both oxidic and sulphidic ores (Rovira & Ambert
2002a, 2002b; Bourgarit & Mille 2005; Bourgarit et al. 2003; Bourgarit 2007) with
the probability that co-smelting occurred.
The picture in Iberia is more complicated. There is fragmentary evidence of copper
oxide smelting slag at Cerro Virtud, southeast Spain which has been radiocarbon
dated to the first half of the 5th millennium BC (Delibes & Montero-Ruiz 1997;
Montero-Ruiz et al. 1999; Ruiz Taboada & Montero-Ruiz 1999). However, it is at
least a millennium older than any other evidence of smelting or anything metallurgical
10
in Iberia (e.g. Montero-Ruiz 1994; 2005). It comprises a piece of slag adhering to a
single ceramic sherd that was excavated under rescue conditions from a site disturbed
by mining and it is dated to a layer, rather than by associated organic material or
feature, which the authors maintain was untouched in spite of the circumstances of
excavation. Whilst sites such as El Palomar and Terrera Ventura in southern Spain
have been cited as potential evidence for 4th millennium BC copper smelting
(Montero-Ruiz 2005), the context of the evidence is not secure whilst the potentially
late 4th millennium BC at the sites such as Rotura, Sala 1 and São Bras 1 remains
unanalysed (Monge Soares et al. 1994). It seems probable that copper smelting
occurred in southern Iberia during the late 4th millennium – early 3rd millennium BC
and is confirmed throughout much of Iberia during the early- mid 3rd millennium BC
at sites such as Bauma del Serrat del Ponte (see Goncalves 1989; Gómez-Ramos
1999; Delibes de Castro & Montero-Ruiz 1999; Fernandez Manzano & Martinez
2003), probably making it contemporary with the earliest gold (Perea 1991; Pingel
1992). Compositional analysis in south-central Portugal has even demonstrated the
recognition and deliberate selection of arsenical copper during the later 3rd
millennium BC, probably through the exploitation of arsenic rich ores (e.g. Müller et
al. 2007). Since Chinflón, southwest Spain has been re-dated to the late 2nd
millennium BC (Pellicer & Hurtado 1980; Rothenberg & Andrews 1996) no copper
mine has been securely dated in Iberia until the early 3rd millennium BC (cf. HuntOrtiz 2003) as at El Aramo, northern Spain (Blas Cortina 2005).
To the north and west of the Alps in central Europe and to the north of the Pyrenees in
Western Europe, there are no metallurgical sources in the geology until Wales,
western and northern Britain and Ireland. In northwest continental Europe, it has been
11
argued that this contributes to the existence of a “Chalcolithic frontier” which
experienced a delayed adoption of metal (e.g. Brodie 1997; 2001). Unsurprisingly,
the evidence for metal production throughout this region is limited with the two
copper droplets found in a hearth at Val-de-Reuil, Seine Valley, dating to the late 3rd
millennium cal BC (Billard et al. 1991) represents a substantial discovery (see
Meurkens 2004). However, the recent re-analysis of early metal objects reveals a far
earlier initial presence than might be expected. The dating of the recently excavated
collective burials at Vignely in north central France of which one, a child of around 5
years old, had a necklace of nine copper beads, produced a range of 3517-3357 cal
BC (Mille & Bouquet 2004) making it older even than the earliest metal objects
throughout southern France (Guilaine 1991; Roussot-Larroque 2005). Similarly,
copper objects are known from the north Alpine region and the northern temperate
regions such as the great European plains and even southern Scandinavia from the late
5th millennium BC (e.g. Ottaway 1973; Klassen 2000; Klassen et al. 2001). By
comparision, the earliest gold objects beyond the Mediterranean are thought to date to
the mid-late 3rd millennium BC (Hartmann 1970; 1982; Elèure 1982; Primas 1995).
There is a broad consensus placing the appearance of metal in the mid 3rd millennium
BC in northwest continental Europe (Cauwe et al. 2001; Warmenbol 2004) though
typologically earlier objects have been recovered as far west as Brittany (Briard &
Roussot-Larroque 2002). Moving across the Channel, surveys and excavations of the
copper ore sources in Wales have revealed extraction activities beginning in earnest c.
2100/2000 cal BC onwards (Timberlake 2002; 2003). The copper sulphide ore
extraction and possibly smelting in southwest Ireland at Ross Island c. 2400 BC
onwards (O’Brien 2004) therefore probably represents the earliest copper production
12
in northwest Europe. The only other broadly contemporary evidence for metal
production consists of a splash of arsenical copper found in a midden at Northton on
the Isle of Harris in Scotland (Brodie 1997) that is dated to the late 3rd millennium BC.
Whilst there is typologically dated evidence for copper or gold objects in Ireland
dating to the earliest period of metal production beyond the metal axe marks in the
Corlea 6 wooden trackway in the Irish Midlands dendro-dated to 2259±9 BC
(O’Sullivan 1996), there are several later 3rd millennium cal BC radiocarbon dates in
southern Britain for copper and gold objects found in the Bell Beaker burial sites of
Barrow Hills, Shrewton, Barnack, Chilbolton and Amesbury (see Needham 1996 for
review; Fitzpatrick 2002).
The appearance and adoption of tin-bronze alloys throughout Western Europe and
beyond appears to have been a sporadic process. Individual objects made of copper
with low additions of tin appear in Montenegro, central Germany, northern Italy and
northern Spain during the earlier 3rd millennium BC (Müller 2003 Krause 2003, 210;
Fernandez Miranda et al. 1995; Primas 2002) but only reaching southern Iberia during
the early 2nd millennium BC (Montero 1994). Whilst the ability to create tin bronze
can be demonstrated at an early date and appears to be transmitted from east to west,
the adoption and production of bronzes with consistently high percentages of tin only
occurred significantly later throughout many regions of Europe (Pare 2000). In
contrast to this is the very rapid transition from the use of pure copper or arsenical
copper to tin bronze in Britain and Ireland during the late 3rd millennium BC
(Needham 1996; O’Brien 2004).
How did Metallurgical Transmission occur?
13
The relatively abundant presence of colourful outcrops of copper ores throughout
Iberia, southeast France, Wales and southern Ireland would seem to imply that the
earliest prospecting in Western Europe would have presented few challenges (Figure
1a-d). However, the presence of other brightly-coloured mineral sources and the
diversity of the copper ore colours could have been a source of confusion for
inexperienced prospectors. Whilst it is highly likely that prehistoric communities
would have observed copper ores during the pre-metallurgy period, they would also
have observed many other mineral sources. As there is no evidence of copper ores or
native copper being exploited during the pre-metallurgical period as occurs in
southeast Europe, there is no sense that a distinction of copper-bearing minerals had
been made or their having any identifiable significance until being recognised for
their metallurgical properties. If consideration is given to the differing requirements
for smelting oxidic and sulphidic copper ores, then not only copper ores have to be
identified but also the copper ores that can be smelted employing existing practices.
Due to the variation in regional geologies, environments and in the accessibility of
copper ores, the discovery of new sources requires flexible prospecting techniques as
well as expectations. The practice of prospecting would therefore have been partially
a process of building on accumulated knowledge of the local landscape and partially
on the experience that enabled the correct identification of an ore source and an ore
type. It is also probable that, rather than envisaging a systematic landscape survey, the
single discovery of a copper, gold, lead or tin source would lead to intensive
prospection of the surrounding area for further potential sources.
14
When compared to later mining techniques (e.g. Craddock 1995), the evidence for
surface and sub-surface extraction of oxidic and sulphidic copper ores1 in Western
Europe at sites such as Les Neuf-Bouches, Cabrières, southeast France from c. 3100
BC (Bouquet et al. 2006), Ross Island, southwest Ireland from c. 2400 BC (O’Brien
2004) and Copa Hill, central Wales from c. 2100/2000 cal BC (Timberlake 2002;
2003) appears simple and straightforward 2 . The presence of earlier and more
extensive underground mines, such as the variscite mines at Can Tintorer, northern
Spain (Blasco et al. 1998; Bosch 2005) dating to the 5th millennium BC or the flint
mines at Cissbury or Grimes Graves in southern England dating to the 4th millennium
BC (Barber et al. 1999), implies that copper extraction represented an adaptation of
earlier practices. However, this does not mean that anyone seeking to extract copper
ore innately possessed the necessary expertise to do so. Ore veins were followed
underground at copper mining sites such as El Aramo, northern Spain (Blas Cortina
2005), Pioch Farrus IV, Cabrières, southeast France (Ambert 1995) and Ross Island,
southwest Ireland (O’Brien 2004) despite the greater need for labour as well as the
organisation and expertise needed to facilitate the movement of miners and their
equipment, in providing adequate ventilation, illumination and drainage, and bringing
the ore to the surface, all whilst ensuring that the underground structures did not
collapse. Opening new sources nearby would have required substantially less effort
and expertise but there is little evidence to suggest that simply any ores or ore sources
were exploited randomly. For each mining site the range of radiocarbon dates indicate
a long-term commitment over centuries though the evidence of inhabitation in the
1
Unfortunately, the evidence for gold, lead and tin extraction is non-existent (cf. Weisgerber &
Pernicka 1995; Meredith 1998), probably because they derived from placer deposits in streams and
surface collection, both of which will have left very ephemeral evidence.
2
Compositional and lead isotope analysis has traced metalwork assemblages to potential mines that
remain currently undated as in Mocissos and Tharsis in southwest Iberia (Nocete 2004; Müller et al.
2007), Herrerías in southeast Spain (Delibes et al. 1989; Montero 1994, 223).
15
immediate environs is limited. For ore extraction to occur, there are the tools and
equipment such as stone hammers and antler picks (e.g. Pascale 2003; Timberlake
2003) that would have to be made and carried, the food that would have to be
acquired, the fuel that would have to be sourced for food but also larger quantities for
fire-setting (cf. Weisgerber and Willies 2001) and people who would have to be
organised, even if rich ore sources were nearby as in southern Iberia (Montero-Ruiz
1994; Hunt-Ortiz 2003). It would not have been possible to procure everything in the
close vicinity of any mining sites, whether due to the stone resources, food production
or the distance from inhabitation sites. The implication is that there would have to
have been dedicated mining expeditions that had access to the ore. This would have
required the active participation of a small group or community containing several
individuals having relevant expertise. Their involvement would be needed in the
processing or beneficiation of the ore whether close to the mining site as at Roque
Fenestre and Farrus 448, southeast France (Ambert et al. 1984; Espérou et al. 1994) or
within inhabitation areas further away at Cabezo Juré, southwest Spain (Nocete 2004)
as well as in its subsequent movement. There are ongoing debates with certain
scholars proposing the control of copper ore sources and production by specific
groups (e.g. Nocete 2001, 2004) as well as those arguing that the abundance of copper
ore and relatively widespread evidence of copper working means that this would not
be possible (e.g. Suarez et al. 1986; Rovira 2002).
All the analysed smelting evidence can be characterised from a technological
perspective as closely resembling each other - small scale, relatively low temperature
processes carried out under poorly reducing conditions on oxidic and/or sulphidic ores
in small stone and clay structures and/or ceramic crucibles with no intentional
16
addition of fluxes and little consequent slag (Craddock 1999; Bourgarit 2007) 3 . The
smelting would yield only small quantities of copper that would then have to be
refined in a separate process. There are several sites where refining rather than
smelting has been identified, but there remain many that are assumed to have smelting
evidence that require analysis. There is no evidence for any fundamental changes in
the smelting practices during the 4th and 3rd millennia BC in Western Europe.
Whether a transfer of existing pyrotechnological capabilities during the premetallurgical period to the smelting of metal can be envisaged depends on the
characteristics of the processes involved. The extensive presence of ceramics
throughout Western Europe before metallurgy provides the most promising evidence
though there are no known ceramic firing sites. It is highly probable that pottery, as
elsewhere in Europe during this time, was fired in an open bonfire, which would
render the process virtually invisible archaeologically (see Orton et al. 1997, 127130). It is therefore the experimental reconstructions of ceramic open bonfire firings
that provide the clearest indications of the pyrotechnological abilities (e.g. Gosselain
1992; Livingstone-Smith 2001; McDonnell 2001). Characteristics of this open firing
technique are a lack of control, rapid changes in temperature, an oxidizing atmosphere
and a duration varying from several minutes to several hours. Though temperatures of
c. 1000ºC can occasionally be reached, this is only for a very short duration and
cannot be maintained before dropping back to 600-800ºC or lower. This failure to
sustain a sufficient temperature that is comparable to experimental reconstructions of
3
The proposal by the excavator that Cabezo Juré, southwest Spain possessed copper smelting furnaces
dating from the early 3rd millennium BC is unconvincing from the published data (Saez et al. 2003;
Nocete et al. 2004; Nocete 2006). It is comparable to metallurgical features at other sites in southern
Iberia which have been demonstrated not to be furnaces. Furthermore, if true these furnaces would be
contemporary with the earliest use in the east Mediterranean and Near East (Hauptmann 2000;
Craddock 2001) and substantially pre-date those furnaces in the central Mediterranean and Alpine
regions that appeared during the later 2nd millennium BC (Craddock 1999).
17
smelting of oxidic and/or sulpidic ores based on evidence and/or probable conditions
in southern Iberia (Rovira and Guttierez 2005), southeast France (Bourgarit et al.
2003), Wales (Timberlake 2005; 2007) and southwest Ireland (O’Brien 2004), the
oxidizing rather than reducing atmosphere, and the lack of control over both makes it
unlikely that copper smelting using a ceramic open firing method could have
occurred.
Furthermore, it is unlikely that charcoal was used in the firing of ceramics or whether
its recognition and use occurred before metallurgy. Wood, peat and dung would have
been available and perfectly sufficient in the creation of pottery and other activities
involving heat and fire and there does not appear to have been any pyrotechnological
reason for charcoal to be employed. For metallurgy, the use of charcoal is particularly
important, not simply due to its ability to create high temperatures using relatively
small quantities in a small space, but due to it being a source of highly-reducing
carbon monoxide gas (see Horne 1982; Craddock 2001). This makes it ideal for
smelting copper ores and its absence renders the smelting process far less effective if
not completely impossible. If any transfer of pyrotechnological knowledge did occur,
it could not have been straightforward or simple. If independent experimentation led
to the successful smelting of copper ores then there would have to have been
significant alterations to the existing practices as well as the independent motivation
to attempt such experimentation in the first place. Neither does it appear that copper
smelting could have been consistently achieved with only a partial knowledge of the
process involved – such as the ability to identify copper ores (as well as differentiate
between oxidic and sulphidic ores), or the need for high temperatures. As modern
experiments have shown, smelting needed to be carried out within a fairly narrow
18
margin of error or else the entire process would fail. Though lead could be smelted at
lower temperatures, there is no evidence that it preceded the smelting of copper in
Western Europe. Similarly, there is no evidence for gold objects preceding copper
despite gold melting being a more straightforward process, albeit at a comparable
temperature to copper smelting. It is argued that for smelting technology to spread,
the expertise would have to be learnt in one place and applied elsewhere. This could
therefore apparently only occur either through the movement of individuals or groups
possessing the smelting skills.
Though there are few smelting sites that have been comprehensively and carefully
excavated and analysed, where this has occurred there is evidence that smelting
continued to be practiced for centuries in the same place whether at La Capitelle du
Broum, southeast France (Ambert et al. 2002; 2005), Cabezo Juré, southwest Spain
(Nocete 2004) or more tentatively at Ross Island, southwest Ireland (O’Brien 2004).
The nature of these places varies - in the southern Iberian regions copper smelting has
been found in the large walled settlements such as Los Millares (Arribas et al. 1989;
Molina et al. 2004), smaller fortified sites such as Cabezo Juré (Nocete 2004; Nocete
2006) and unfortified sites such as Almizaraque (Delibes et al. 1986; 1989; Müller et
al. 2004). This diversity is also seen in southeast France with the excavation of the
extensive double-walled structures at the “metallurgical village” of La Capitelle du
Broum (Ambert et al. 2002; 2005) in contrast with settlement sites such as Al Claus
(Carozza et al. 1997; Carozza 1998) with less substantial structures. Though there is
only a single known potential smelting site in Ross Island, southwest Ireland, its
temporary huts stand in contrast to the sites in the other regions (O’Brien 2004).
These places where copper smelting occurred, whether in the earliest form or
19
centuries later, appear to be located close to the extraction and ore processing sites,
generally close to the available ore sources. The surviving evidence of the smelting
equipment comprises thick-walled open-mouthed ceramic vessels as in Iberia and
southeast France (Rovira & Ambert 2002a; 2002b) as well as clay lined hearths in
southern Iberia as famously excavated at Los Millares (Hook et al. 1989; 1991)
(Figure 2) and Zambujal (Müller et al. 2007). The virtual absence of such structures in
Ireland and Britain can be explained by the potentially archaeologically ephemeral or
even invisible evidence from smelting (cf. Timberlake 2005; 2007). The virtual
absence of
ceramic or stone tuyère fragments throughout Western Europe and
beyond during the 4th and 3rd millennia BC (Roden 1988) has led to suggestions of
wind potentially playing a greater role than previously acknowledged (e.g. Happ
2005; Nocete 2004; Nocete 2006; Bourgarit 2007, 7-8).
Arsenical copper is probably the earliest intentional copper alloy appearing
throughout Western Europe from at least the 3rd millennium BC. Its origins are hard
to assess as some of these are intentionally-produced alloys and others were probably
the result of accidental smelting of copper ores with arsenic impurities or mixed ore
charges. There does not appear to be any systematic shift from producing copper
metal to copper arsenic alloys. Rather, in certain instances, there appears to be the
opportunistic exploitation of copper arsenic metals due to the smelting of copper ores
rich in arsenic that may have been accompanied by an awareness of how the metal
could be reproduced. This translates into the creation of specific object forms such as
elongated awls and sheet metal at Zambujal south-central Portugal (e.g. Müller et al.
2007) and possibly halberds and daggers in Ireland (Northover 1989) as well as the
contemporary presence but less clearly deliberate exploitation of the alloy as in
20
southern Iberia (Montero-Ruiz 1994, 247-263; Hunt-Ortiz 2003 contra Hook et al.
1991; Keesman et al. 1991/2). There is no evidence that gold, copper or lead were
mixed to form alloys. These widely differing rates of adoption of tin bronze do not
appear to be related to the distance away from the relatively scarce tin ores throughout
Europe (Pernicka 1998). Rather, it appears to relate to regional preferences (see
Primas 2002; Guimlia-Mair & Lo Schiavo 2003) even though the smelting of tin ore
or the direct creation of copper-tin through mixed ore smelting would have been
within the capabilities of a copper smelter.
The realisation that the metal can be cold-worked for a longer time if heated in
between shaping will not have escaped the notice of people who were used to fireharden wood, heating flint and firing pottery and it is therefore no surprise to find it
present throughout Western Europe in the earliest objects. However, the application
of heat to create a liquid from a solid that could then be poured into a mould to form a
new object when cooled, does not have real parallels in pre-metallurgical societies but
are still present in the vast majority of the earliest copper objects in Western Europe
(e.g. Rovira & Gómez-Ramos 2004). The excavation of equipment relating to the
creation of copper and gold objects, such as moulds, hammers, tongs and anvils, is
very sparse relative to the number of objects that have been recovered. This is
partially due to the difficulty in identifying the specific tools that would have been
employed, but more probably related to the rapid degradation of sand moulds (e.g.
Ottaway & Seibel 1997; Eccleston & Ottaway 2002), the fragmentation of clay
moulds (e.g. Ottaway 2003) and the decomposition of any wooden objects such as
patterns, models and containers.
21
Where exhaustive typological research has been conducted on an object type such as
beads and copper flat axes in southeast France (Chardenoux & Courtois 1979; Barge
1982) and flat axes, halberds and daggers in Ireland (Harbison 1969a; 1969b), it has
revealed extensive morphological micro-variations based on several distinctive
designs. Despite the possibility of specific object designs being reproduced in copper
and gold, there is little evidence for any imitation beyond objects designed to worn as
personal adornment (e.g. Barge 1982). It appears that the replication of specific
objects occurred far less frequently than the creation of subtly new ones. The
implication for the production of copper objects is that slight alterations on accepted
norms occurred. Rather than re-use stone moulds or wooden patterns for shaping clay
and sand moulds, new moulds would have to be made or the metal would have to be
manipulated in a different way.
The quantity of metal objects produced is always going to be far greater than the
number recovered. It is through the deposition or discarding of copper and gold
objects in each region, rather than their recycling or re-melting, that the perception of
the presence of metal is those prehistoric communities is shaped. Consequently, there
is an unavoidable bias in the perception of metal use in the past towards regions or
objects where higher metal deposition, (rather than recycling), occurred (cf. Needham
1998; 2001; Taylor 1999). However, there is little indication of large scale production
even in regions surrounding mining and smelting sites. For instance, in southeast
France there are only 1403 known copper objects dating to the later 4th millennium –
mid 3rd millennium which is still double that of southeast Spain during the same
period (Montero 1994) (Figure 3). The object forms involved range in diversity and
emphasis with copper flat axes, beads, needles, fishhooks, awls, knives, daggers,
22
saws, sickles, spatulas, chisels found in Iberia (Delibes de Castro & Montero-Ruiz
1999). In contrast, there is a more restricted range with a distinct emphasis on copper
beads in southeast France (Gutherz & Jallot 2005) and a virtual absence of any copper
objects beyond flat axes, daggers and halberds in Ireland (Harbison 1969a; 1969b).
The life of metal objects after they are produced and before they are discarded or
deposited is the most elusive part of their existence. Aspects such as where an object
was taken, how it was used, how it changed possession, the perceptions that
surrounded it and whether it was recycled, re-melted or re-cast may well be more
important to understanding its presence than production or depositional practices.
However, approaching any of these aspects is difficult using current archaeological
and archaeometallurgical techniques. With a chronological resolution of centuries, it
is the identification of broader patterns of metal movement, use and circulation
relative to metal production sites that can be investigated. The extensive programmes
of compositional and latterly lead isotope analyses of early copper objects in Western
Europe reveals the circulation of metal objects from production centres. For instance,
the vast majority of copper objects in Ireland and Western Britain conform to high
arsenic and antimony, as well as a lead isotope, signatures that in all probability
derives from Ross Island, southwest Ireland (Coghlan & Case 1957; Case 1966; Rohl
& Needham 1998; Northover et al. 2001). This exploitation of a single source for
several centuries by communities in a region can also be seen in southeast France
when the high silver and antimony compositions and lead isotope levels were
examined against the ores at Cabrières (Ambert 1999; Prange et al. 2003; Prange &
Ambert 2005). In contrast, the recent re-analyses of compositional data in eastern
Britain and continental northwest Europe reveals very different yet coherent
23
patterning, originally termed Bell Beaker metal (Butler & Van der Waals 1964; Butler
& Waterbolk 1965), that appears to be formed from the mixing of metals from two
obviously distant geological sources, with one perhaps even as far as northern Spain
(Needham 2002). To investigate the archaeological contexts in which metal objects
are found requires a high standard of excavation, recording and publication. As the
vast majority of the metal objects in each region were accidentally discovered during
the 19th to mid 20th century, the quality of the data tends to be highly variable. At best,
it is possible to discern broad patterns that can be contrasted such as the
concentrations of metal ornaments in burial contexts accompanying bodies as
compared to the concentrations of flat axes placed elsewhere (e.g. Needham 1988;
Gutherz & Jallot 2005).
Metals, Metallurgy and Material Traditions
It can be demonstrated that expertise would have to have been gained for certain
aspects of metal production in Western Europe, such as the selection and smelting of
the ore, for a successful transfer from individual to individual making it probable that
metallurgical skills were restricted, whether intentionally or not. The inevitable or
deliberate restriction of such crucial knowledge such as the correct raw materials, the
smelting equipment or the sequence and timing of actions and addition of substances
could have ensured that it remained in the hands of a few select groups of metal
producers, who only passed on their craft to people of their choosing. If the
ethnographic record is any guide, in virtually all instances this means specific
members of an extended family or tribe with songs, rituals and taboos reinforcing the
restricted knowledge and expertise. It is argued here that the requirement of a
24
metallurgical apprenticeship (cf. Keller 2001) and the subsequent movement of
metallurgists, either returning to their original regions or settling in new ones, would
create extensive yet fragile networks of expertise.
However, it is important not to over-emphasise the primacy of metal production
techniques in this process. It is argued that the desires of the communities supporting
metal production as metal consumers were actually more important. For metal
orientated networks of production expertise to exist, individuals and communities are
needed to invest in the acquisition of metal objects and metallurgical technology. It is
even perhaps erroneous to discuss individuals in certain aspects of the metallurgy
given the collective nature of so many of the production processes, including the ore
prospection, extraction, processing and transport. Even the existence of part-time
copper smelters or smiths, who are more likely to have had distinctive individual roles
owing to their specialist expertise, required the commitment of the broader
community to aid in the procurement of food and shelter as well as production and
acquisition of the objects. There is no inherent functional reason why metal objects or
metal production should be adopted by local communities or introduced by non-local
communities. It is important to re-emphasize that the earliest copper, gold and lead
objects were not necessarily superior to wood, bone, flint and ceramics for performing
everyday tasks and that there were many obstacles and complications involved in
metal production practices relative to those in existing materials. The distinctive
colours, lustre, malleability and ability to carry decorations and be recycled can be
proposed as attractive qualities for adopting metal (cf. Keates 2002). However, early
metal objects and production practices were adopted due to specific community
desires and standards (see Helms 1993; Sofaer Derevenski & Sørensen 2002; in press)
25
which created and reproduced spatially and temporally distinct metallurgical
traditions throughout Western Europe (Roberts in press b).
Metallurgical traditions did not exist independently and were closely bound into other
larger and more pervasive socio-cultural networks and traditions whose spatial and
temporal variations are expressed through metal as well as other material culture (e.g.
Sofaer 2006). For instance, the influence of the Bell Beaker phenomenon during the
mid third millennium BC in creating and shaping existing copper metallurgical
traditions varies throughout Western Europe (see Brodie 1997; 2001; Vander Linden
2006; 2007). In southeast France, there is a marked reduction in the diversity and
quantity of metal objects (Ambert 2001), in Iberia new forms are simply added to the
existing repertoire without any discernable change in the underlying technology
(Rovira 1998; Rovira and Gomez-Ramos 2004), in northwest continental Europe and
southern England communities are seemingly obtaining metal objects from two
distant sources which are then melted, mixed together and re-cast (Needham 2002). In
Ireland, despite the chronological parallels in the appearance of metal and Bell Beaker
pottery and the discovery of several sherds at the metal production site of Ross Island,
southwest Ireland (O’Brien 2004), metal appears to re-define existing traditions rather
than adorn the dead in new burial rites as in southern England (Needham 1996;
Fitzpatrick 2002). The extraction of the raw material for producing copper axes at a
specific place, the widespread distribution of the flat axes, and the virtual absence of
these axes in burial contexts parallels earlier traditions of polished stone axes in
Ireland (e.g. Cooney and Mandal 1998). It is possible to observe the replacement of
polished stone axes by metal flat axes throughout Ireland, demonstrated by the
subsequent decrease of polished stone axes during the mid-late third millennium BC,
26
a process that is especially acute in southwest Ireland (O’Brien 2004, 562). It is
evidence that despite the many possibilities afforded by the new material and new
inter-connections generated by the technology, there were evidently boundaries in
form and composition that could not be transgressed. The ability to recycle metal
means that object forms created elsewhere could be melted down and converted into
more familiar shapes, even in regions far from copper ore deposits. It is this potential
fluidity of a widely recognised material that would have been important in the
sustaining and intensifying metallurgical networks and traditions (cf. Shennan 1993;
1999; Sherratt 1974; 1993). Despite the potential and possibilities, metal in the fourth
and third millennia BC is relatively small scale and its development is very gradual.
Neither the production nor the consumption of metal possesses serious enough
credentials to be considered a major, let alone revolutionary, influence in the broader
worlds of the communities involved (Chapman 2003; Bartleheim 2007).
Acknowledgements
This paper arises out of my doctoral research supervised by Marie-Louise StigSørensen and Barbara Ottaway and funded by the Domestic Research Studentship.
Though I have benefited immensely from conversations with many scholars, the
errors and opinions remain my own.
Bibliography
Ambert P. 1995. Les mines préhistoriques de Cabrières (Hérault) : quinze ans de
recherche. Etat de la question. Bulletin de la Société Préhistorique Française 92 (4),
499-508.
Ambert, P. 1999. Les minerais de cuivre et les objets metalliques en cuivre à
antimoine-argent du sud de la France. Preuves d'une exploitation minière et
métallurgie du début du III millenaire av. J.C. In A. Hauptmann, E. Pernicka,T.
Rehren & Ü. Yalçin (eds.). The Beginnings of Metallurgy, pp 176-192. Bochum: Der
Anschnitt.
Ambert, P. 2001. La place de la métallurgie campaniforme dans la première
27
métallurgie française. Etat de la question. In F. Nicolis (ed.) Bell Beakers Today:
Pottery, people, culture and symbols in prehistoric Europe, pp. 387-409. International
Colloquium Riva del Garda (Trento, Italy), 11-16 May 1998.Trento: Ufficio Beni
Culturali.
Ambert, P, Barge, H., Bourhis, J.R. & J.L. Esperou, 1984. Mines de cuivre
préhistorique de Cabrières (Hérault): Premiers resultants. Bulletin de la Société
Préhistorique Française 81 (3), 83-89.
Ambert, P., Bouquet, L., Guendon, J.L. and Mischka, D. 2005. La Capitelle du Broum
(district minier de Cabrières-Péret, Hérault): établissement industriel de l’aurore de la
métallurgie française (3100-2400 BC). In P.Ambert and J. Vaquer (eds) La première
métallurgie en France et dans les pays limitrophes, pp 83-96. Mémoire de la Société
Préhistorique Française 37.
Ambert, P. & Carozza, L. 1998. Origine(s) et dévelopment de la première
métallurgie française. Etat de la question. In B. Fritsch, M. Maute, I. Matuschik, J.
Müller & C. Wolf (eds). Tradition und Innovation. Prähistorische Archäologie als
historische Wissenschaft. Festschrift für Christian Strahm, pp 149-173. Rhaden:
Verlag Marie Leidorf.
Ambert, P., J. Coularou, C. Cert, J-L. Guendon, D. Bourgarit, B. Mille, N. Houlès and
B. Baumes, 2002 Le plus vieil établissement de metallurgistes de France (III
millénaire av J.-C.): Péret (Hérault). Comptes Rendu Palevol 1: 67-74.
Arribas, A., Craddock, P. Molina, F, Rothenberg, B & Hook, D. 1989. Investigacion
arqueo-metalurgia en yacimientos de los edades de cobre y del bronce en el sudeste de
Iberia. In C Domergue (ed.) Minería y Metalurgia de las Antiguas Civilizaciones
Mediterráneas y Europeas, pp 71-79. Madrid: Ministerio de Cultura.
Bailey, D.W. 2000. Balkan Prehistory. London: Routledge.
Barber, M., D. Field, and P. Topping 1999. The Neolithic Flint Mines of England.
Swindon: English Heritage
Barfield, L. 1966. Excavations on the Rocca di Rivoli (Verona) 1963 and the
prehistoric sequence in the Rivoli basin. Memorie del Museo Civico di Storia
Naturale, Verona 14: 1-100.
Barge, H. 1982. Les Parures du Neolithique Ancien au debut de l’Age des Metaux en
Languedoc. Paris: CNRS.
Bartelheim, M. 2007. Die Rolle der Metallurgie in vorgeschichtlichen Gesellschaften:
Sozioökonomische und kulturhistorische Aspekte der Ressourcennutzung. Ein
Vergleich zwischen Andalusien, Zypern und dem Nordalpenraum. The role of
metallurgy in prehistoric societies: Socioeconomic and cultural aspects of the use of
resources. A comparison between Andalusia, Cyprus and the north Alpine area.
Forschungen zur Archäometrie und Altertumswissenschaft 2. Rahden: 2007.
28
Bartleheim, M. Eckstein, K, Huijsmans, M., Krauss, R. & Pernicka, E. 2002.
Kupferzeitliche Metallgewinnung in Brixlegg, Österreich. In M. Bartelheim, E.
Pernicka & R. Krause (ed.) The Beginnings of Metallurgy in the Old World, pp 33-82.
Rahden: Verlag Marie Leidorf.
Billard, C., Bourhis, J.-R., Desfosses, Y., Evin J., Huault, M., Lefebvre, D., PauletLocard, M.-A. 1991. L'habitat des Florentins à Val-de-Reuil (Eure). Gallia préhistoire,
33, 140-171.
Blasco, Anna, Villalba, María Josefa, Edo, Manuel 1998 Explotación. manufactura,
distribución y uso como bien de prestigio de la "calaita" en el Neolítico: el ejemplo
del complejo de Can Tintorer. In Minerales y metales en la prehistoria reciente:
algunos testimonios de su explotación y laboreo en la Península Ibérica, Germán
Delibes de Castro ed. pp 41-70. Studia Arqueológica 88. Valladolid. Universidad de
Valladolid.
Blas-Cortina, M., 2005. Un témoignage probant de l’exploitation préhistorique du
cuivre dans le nord de la Péninsule Ibérique: le complexe minier d’El Aramo
(Asturias). In P.Ambert and J. Vaquer (eds) La première métallurgie en France et
dans les pays limitrophes, pp 195-206. Mémoire de la Société Préhistorique Française
37.
Bognar-Kutzian, I. 1976. On the origins of early copper-processing in Europe. In
Megaw, J.V.S., (ed.) To Illustrate the Monuments, pp. 69-76, London: Thames &
Hudson.
Bosch, J. 2005. Les techniques d’exploitation des plus anciennes mines d’Europe
méditerranéenes: l’example de Gavá, Barcelona. In P.Ambert and J. Vaquer (eds) La
première métallurgie en France et dans les pays limitrophes, pp 207-210. Mémoire de
la Société Préhistorique Française 37.
Bourgarit, D., 2007. Chalcolithic Copper Smelting. In Susan La Niece, Duncan Hook
and Paul Craddock eds. Metals and Mines: studies in archaeometallurgy pp 3-14.
London: Archetype and British Museum
Bourgarit, D. & Mille, B. 2005. Les nouvelles données de l’atelier metallurgique
chalcolithique de La Capitelle du Broum dans le district de Cabrières (Hérault): la
transformation des minerais de cuivre à base de sulfures se précise. In P. Ambert & J.
Vaquer (eds) La première métallurgie en France et dans les pays limitrophes, pp 97108. Mémoire de la Société Préhistorique Française 37.
Bourgarit, D., Mille, B., Prange, M., Hauptmann. A. and Ambert, P. 2003.
Chalcolithic fahlore smelting at Cabrières: reconstruction of smelting processes by
archaeometallurgical finds. In Archaeometallurgy in Europe pp 431-440. Milan:
Associazione Italiana di Metallurgia
Bouquet L., Figueroa-Larre V., Laroche M., Guendon, J-L. and Ambert, P. 2006. Les
Neuf-Bouches (district minier de Cabrières-Péret), la plus ancienne exploitation
29
minière de cuivre de France: travaux récents, consequences. Bulletin de la Société
Préhistorique Française 103 (1), 143-159.
Briard, J. & Roussot-Larroque, J. 2002. Les Debuts de la Metallurgie dans la France
Atlantique. In M. Bartelheim, E. Pernicka & R. Krause (ed.) The Beginnings of
Metallurgy in the Old World, pp 135-160. Rahden: Verlag Marie Leidorf.
Brodie, N. 1997. New perspectives on the Bell Beaker Culture. Oxford Journal of
Archaeology 16, 297-314.
Brodie, N. 2001. Technological frontiers and the emergence of the Beaker culture. In .
Nicolis (ed.) Bell Beakers Today: Pottery, people, culture and symbols in prehistoric
Europe, pp. 487-496. International Colloquium Riva del Garda (Trento, Italy), 11-16
May 1998.Trento: Ufficio Beni Culturali.
Budd, P. and T. Taylor. 1995. The faerie smith meets the bronze industry: magic
versus science in the interpretation of prehistoric metal-making. World Archaeology
27, 133-43.
Butler, J.J. & J.D. Van der Waals. 1964. Metal Analysis: SAM 1 and European
Prehistory. Helenium 4, 3-39.
Camps, G. 1988. Terrina et le Terrinien: recherches sur le Chalcolithique de la
Corse. Rome: Ecole Française de Rome 109.
Carozza, L., Bourgarit, D, Mille, B., & Burens, A. 1997. L’habitat et l’atelier de
metallurgiste chalcolithique d’Al Claus: analyse et interprétation des témoins d’
activité metallurgique. Archéologie en Languedoc 21 147-164.
Cartihallac, E. 1886. Les ages préhistorique de l’Espagne et du Portugal. Paris: C.
Reinwald.
Case, H. 1966. Were Beaker people the first metallurgists in Ireland? Palaeohistoria
12, 141-177.
Cauwe N., Vander Linden M. & Vanmontfort B., 2001. The Middle and Late
Neolithic. In Cauwe N, Hauzeur A & van Berg P.-L. (eds) Prehistory in Belgium –
Préhistoire en Belgique. Bruxelles: Société Royale Belge d’Anthropologie et de
Préhistoire (Anthropologica et Praehistorica 112, n° spécial à l’occasion du XIVe
congrès de l’Union Internationale des Sciences Pré- et Protohistoriques) : 77-89.
Chapman, R.W. 2003. Archaeologies of Complexity. London: Routledge.
Chardenoux, M-B & Courtois, J. 1979 Les haches dans la France méridonale.
Praehistorische Bronzefunde. Abteilung IX, II Munich
Chernykh, E.N. 1978. Ai Bunar, a Balkan copper mine of the IVth millennium BC.
Proceedings of the Prehistoric Society 44: 203-218.
30
Childe, V.G. 1930. The Bronze Age. Cambridge: Cambridge University Press.
Childe, V. G. 1944. Archaeological ages as technological stages. Journal of the Royal
Anthropological Institute of Great Britain and Ireland 74: 7-24.
Coghlan, H.H. & H. Case. 1957. Early metallurgy in Ireland and Britain. Proceedings
of the Prehistoric Society 23, 91-123.
Cooney, G. and Mandal, S. 1998. Irish Stone Axe Project I. Dublin
Craddock, P. 1995. Early Mining and Metal Production. Edinburgh: Edinburgh
University Press.
Craddock, P.1999. Paradigms of metallurgical innovation in prehistoric Europe. In A.
Hauptmann, E. Pernicka, T. Rehren & Ü. Yalçin (eds) The Beginnings of Metallurgy,
pp 175-192. Der Anschnitt Beiheft 9. Bochum: Bergbau-Museum.
Craddock, P., 2001 From hearth to furnace: evidences for the earliest metal
smelting technologies in the eastern Mediterranean. Palaeorient 26: 151-165.
Delibes de Castro, G., M. Fernández-Miranda, M.D. Fernández-Posse, C. Martin, S.
Rovira & Sanz, M. 1989. Almizarque (Almería): minería y metalurgia calcolíticas en
el sureste de la Peninsula Ibérica. In C Domergue (ed.) Minería y Metalurgia de las
Antiguas Civilizaciones Mediterráneas y Europeas, pp 81-96. Madrid: Ministerio de
Cultura.
Delibes de Castro, G. and I. Montero. 1999. Las Primeras Etapas Metalurgicas en la
Peninsula Ibérica. II. Estudios Regionales. Madrid: Instituto Universitario Ortega y
Gasset y Ministerio de Educacion, Cultura y Deporte.
Domergue, C. 1987. Catalogue des mines et des fonderies antiques de la Penínsule
Ibérique. Madrid: Publications de la Casa de Velázquez.
Eccelston, M & Ottaway, B.S. 2002. Experimental casting of copper and bronze in
sand moulds. In E Jerem & K.T. Biro (eds.) Archaeometry 1998. Oxford: British
Archaeological Reports (International Series) 1043.
Ehrhardt, K.L. 2005. European Metals In Native Hands: Rethinking The Dynamics of
Technological Change 1640-1683. Tuscaloosa: University of Alabama Press.
Eluère, C. 1982. Les Ors Préhistoriques. Paris.
Eogan, G., 1994. The Accomplished Art: gold and goldworking in Britain and Ireland
during the Bronze Age (c. 2300-650 BC). Oxford: Oxbow Monograph 42.
Ehrhardt, K.L. 2005. European Metals In Native Hands: Rethinking The Dynamics of
Technological Change 1640-1683. Tuscaloosa: University of Alabama Press.
Espérou, J.L. 1993. La Structure Métallurgique Chalcolithique de Roque Fenestre à
31
Cabrières. Archéologie en Languedoc 17, 32-46.
Espérou, J.L., Ambert, P., Bourhis, J.R., Roques, P., Gilot, E. et Chabal, L. 1994. La
fosse chalcolithique Pioch-Farrus 448 (Cabrières, Hérault), datation 14C etdocuments
métallurgiques. Bulletin du Musée d'Anthropologie Préhistorique de Monaco 37, 5362.
Fasnacht, W. 1991. Analyses d’objets en cuivre du Néolithique Récent du basin
zurichis. In C. Eluère and J.P. Mohen (eds) Découverte du métal, pp 157-166. Paris:
Picard.
Fasnacht, W. 1995. Metallurgie. In W. Stöckli (ed.) Die Schweiz vom Paläolithikum
bis zum frühen Mittelalte. Volume II Neolithikum, pp183-187. Basel: Verlag
Schweizerische Gesellschaft für Ur-und Frühgeschichte
Fernandez-Manzano, J., and J. I. Martinez. (eds.) 2003. Mineros y Fundidores en el
inicio de la Edad de los metales: el Midi Frances y el norte de la Peninsula Iberica.
Leon: Editorial.
Fernández Miranda, M., Montero-Ruíz, I. and S. Rovira, 1995 Los Primeros Objectos
de Bronce en el occidente del Europa. Trabajos de Prehistoria 52 (1): 57-69.
Fitzpatrick, A.P. 2002. The Amesbury Archer: a Well-Furnished Early Bronze Age
Burial in Southern England. Antiquity 76: 629–630
Fontijn, D., 2002 Sacrificial Landscapes: cultural biographies of persons, objects
and natural places in the Bronze Age of southern Netherlands, c. 2300-600 cal BC.
Annalecta Praehistoria Leidensia 33-34.
Gilman, A. 1996. Craft Specialization in late prehistoric Mediterranean Europe. In B.
Wailes (ed.) Craft Specialization and Cultural Evolution: in memory V. Gordon
Childe, pp 67-71. University of Pennsylvania Museum Monograph 93. Philadelphia:
The University Museum of Archaeology and Anthropology
Giumlia-Mair, Alessandra 2005 Copper and copper alloys in the south-eastern Alps:
an overview. Archaeometry 47 (2): 275-292.
Giumlia-Mair, A. & F. Lo Schiavo (eds.) Le problème de l'étain à l'origine de la
métallurgie / The Problem of Early Tin Colloque / Symposium. Acts of the XIVth
UISPP Congress, University of Liège, Belgium, 2-8 September 2001. British
Archaeological Reports S1199. Oxford: Archaeopress.
Glumac, P.D. 1988. Copper mineral finds from Divostin. In McPherron, A. &
Srejovic, D., (eds.) Divostin and the Neolithic of Central Serbia. pp 457-462.
Pittsburgh: Dept. of Anthropology, University of Pittsburgh.
Glumac, P.D. & J.A. Todd 1991. Early metallurgy in South-East Europe: the evidence
of production. In P.D. Glumac, (ed.), Recent Trends in Archaeometallurgical research,
pp 8-19. Masca Research Papers in Science and Archaeology 8. Philadelphia:
University Museum.
32
Gómez-Ramos, P. 1999. Obtención de los Metales en la Prehistoria de la Península
Ibérica. Oxford: British Archaeological Reports (International Series) 753.
Goncalves, V. 1989. Megalitismo e metalurgia do Alto Algave Oriental: uma
aproximacao integrada. Lisbon: INIC.
Gopher, A., Tsuk, T., Shalev, S. & Gophna, R. (1990) Earliest gold artifacts in
theLevant. Current Anthropology 31.4: 436-443.
Gosden, C. & Y. Marshall (eds.) 1999. The Cultural Biographies of Objects. World
Archaeology 31.
Gosselain, O.P. 1992. Bonfire of the Enquiries. Pottery Firing Temperatures in
Archaeology: What For? Journal of Archaeological Science 19, 243-59.
Guilaine, J. 1991. Roquemengarde et les débuts de la métallurgie en France
méditerranéenne. In C. Eluère and J.P. Mohen (eds) Découverte du métal, pp 279-294.
Paris: Picard.
Guilaine, J. & Elèure, C. 1997. Sur les origines de la métallurgie de l’or dans les
Corbières. In B. Fritsch, M. Maute, I. Matuschik, J. Müller & C. Wolf (eds). Tradition
und Innovation. Prähistorische Archäologie als historische Wissenschaft. Festschrift
für Christian Strahm, pp 175-182. Rhaden: Verlag Marie Leidorf.
Gutherz, X, and Jallot, L. 2005. Âge du cuivre et changements sociaux en Languedoc
méditerranéen. In P.Ambert and J. Vaquer (eds) La première métallurgie en France et
dans les pays limitrophes, pp 83-96. Mémoire de la Société Préhistorique Française
37.
Happ, J., Ambert, P., Bourhis, J.R. & Briard, J. 1994. Premiers essais de métallurgie
expérimentale à l’Archéodrome de Beaune à partir de minerais chalcolithiques de
Cabrières (Hérault). Bulletin de la Société Préhistorique Française 91 (6): 429-434.
Happ, J. 2005. Comment j’ai enterré ma tuyère et mon soufflet à la Capitelle du
Broum. Cu + 29-31.
Harbison P. 1969a. The Daggers and the Halberds of the Early Bronze Age in
Ireland. Munich: Prähistorische Bronzefunde VI,1.
Harbison, P. 1969b. The Axes of Early Bronze Age in Ireland. Munich: Prähistorische
Bronzefunde VI, 2.
Harrison, R.J. 1974. A reconsideration of the Iberian background to Beaker
metallurgy. Palaeohistoria 16, 63-105.
Hartmann, A., 1970 Prähistorische Goldfunde aus Europa: Spektralanalytische
Untersuchungen und deren Auswertung. Studien zu den Anfängen der Metallurgie 3.
Berlin: Mann.
33
Hartmann, A. 1979. Irish and British gold types and their West European
counterparts. In M. Ryan (ed.) The Origins of Metallurgy in Atlantic Europe, 215-228.
Dublin: Stationary Office.
Hartmann, A., 1982 Prähistorische Goldfunde aus Europa: Spektralanalytische
Untersuchungen und deren Auswertung. Studien zu den Anfängen der Metallurgie 5.
Berlin: Mann.
Hauptmann, A. 2000. Zur Frühen Metallurgie des Kupfers in Fenan/Jordanien.
Bochum: De Anschnitt Beiheft 11.
Helms, M.W. 1993. Craft and the Kingly Ideal: Art, Trade and Power. Austin:
University of Texas Press.
Höppner, B., Martin Bartelheim, Melitta Husijmans, Rüdiger Krauss, K. Martinek,
Ernst Pernicka, Roland Schwab, 2005 Prehistoric copper production in the Inn Valley,
Austria, and the earliest copper production in central Europe. Archaeometry 47 (2):
293-315.
Horne, L. 1982. Fuel for the metal worker. The role of charcoal and charcoal
production in ancient metallurgy. Expedition 25, 6-13
Hosler, D. 1995. Sound, colour and meaning in the metallurgy of ancient west Mexico.
World Archaeology 27, 100-115.
Hunt-Ortiz, M.A. 2003. Prehistoric Mining and Metallurgy in South West Iberia
(British Archaeological Reports (International Series) 1188. Oxford: Archaeopress.
Jovanovic B. 1991. La métallurgie énéolithique du cuivre dans les Balkans, et ses
sources en matières premières. In C. Eluère & J.P. Mohen (eds) Découverte du métal,
pp 93-102. Paris: Picard.
Junghans, S., E. Sangmeister & M. Schröder. 1960. Studien zu den Anfängen der
Metallurgie. Berlin: Gebr. Mann
Junghans, S., E. Sangmeister & M. Schröder. 1968. Studien zu den Anfängen der
Metallurgie. Berlin: Gebr. Mann.
Junghans, S., E. Sangmeister & M. Schröder. 1974. Studien zu den Anfängen der
Metallurgie. Berlin: Gebr. Mann.
Kassianidou, V. and Knapp, A.B., 2005 Archaeometallurgy in the Mediterranean: the
social context of mining, technology and trade. In A. Bernard Knapp & E. Blake (eds.)
The Archaeology of Mediterranean Prehistory, pp 215-251. London: Blackwell.
Keates, S. 2002. The Flashing Blade: Copper, Colour and Luminosity in North Italian
Copper Age Society. In A. Jones and G. MacGregor (eds) Colouring the Past: The
Significance of Colour in Archaeological Research pp 109-125. Oxford: Berg.
Keesman, I., Moreno-Onorato, A., Kronz, A. 1991/2. Investigaciones cientificas de la
34
metalurgia de El Malagon y Los Millares, en el Sureste de Espana. Cuadernos de
Prehistoria de la Universidad de Granada 16-17, 247-302.
Keller, C.M. 2001. Thoughts and Production: insights of a practitioner. In M.
Schiffer, (ed.) Anthropological Perspectives on Technology, pp 33-45. Albuquerque:
University of New Mexico Press.
Killick, D. 2001. Science, speculation and the origins of extractive metallurgy. In D.R.
Brothwell and A. M. Pollard, (eds.) Handbook of Archaeological Sciences, pp. 483492. Chichester: John Wiley and Sons.
Kingery, D.W. 1993. Technological Systems and Some Implications with Regard to
Continuity and Change. In S. Lubar and D.W. Kingery (ed.) History From Things
pp.215-230. Washington: Smithsonian Institution Press.
Kingery, D.W. 1996. Introduction. In D.W. Kingery (ed.), Learning from
things: Method and theory of material culture studies, pp 1-15. Washington, DC:
Smithsonian Institution Press.
Klassen, L. 2000. Frühes Kupfer im Norden. Århus: Århus Archaeology Monograph.
Klassen, L., A, Stürup & S. Stürup 2001. Decoding Riesebuch-copper: lead isotope
analysis applied to Early Neolithic copper finds from south Scandinavia.
Praehistorische Zeitschrift 76, 55-73.
Knappett, C. 2005. Thinking Through Material Culture: An Interdisciplinary
Perspective. Philadelphia: University of Pennsylvania Press.
Krause, R. 2003. Studien zur kupfer- und frühbronzezeitlichen Metallurgie zwischen
Karpatenbecken und Ostsee. Leidorf: Rahden/Westf.
Kingery, D.W. 2001. The design process as a critical component of the anthropology
of technology. In M. Schiffer, (ed.) Anthropological Perspectives on Technology, pp
123-138. Albuquerque: University of New Mexico Press.
Lechtman, H., 1977. Style in technology: some early thoughts. In H. Lechtman and R.
Merrill (ed.) Material Culture: Styles, Organization, and Dynamics of Technology, pp
3-20. St. Paul: West Publishing Co.
Lechtman, H. 1996. Cloth and Metal: The Culture of Technology. In E. Boone (ed.)
Andean Art at Dumbarton Oaks, pp. 33-43. Washington DC: Dumbarton Oaks
Research Library and Collection.
Lichardus, J. & Echt, R. 1991. Die Kupferzeit Als Historische Epoche. Symposium
Saarbrucken Und Otzenhausen.
Livingstone-Smith, A. 2001. Bonfire II: The return of Pottery Firing Temperatures.
Journal of Archaeological Science 28, 991-1003.
35
Lo Schiavo, F., 1988 Early Metallurgy in Sardinia. In R. Maddin (ed.) The Beginnings
of the Use of Metals and Alloys. 92-103. Cambridge, MA: MIT Press.
Lo Schiavo, Fulvia, Giumlia-Mair, Alessandra, Valera, Roberto 2005.
Archaeometallurgy in Sardinia: From the Origin to the Beginning of Early Iron Age.
Monographies Instrumentum 30. Montagnac: Monique Mergoil.
Maggi, R. & Pearce, M. 2005. Mid fourth-millennium copper mining in Liguria,
north-west Italy: the earliest known copper mines in Western Europe. Antiquity 79,
66-77.
Makkay, J. 1991. The most ancient gold and silver in Central and South-East Europe.
In C. Eluère & J.P. Mohen (eds) Découverte du métal, pp 119-130. Paris: Picard.
Matuschik, I. 1992. Sengkofen-‘Pfatterbreite’, eine Funstelle der Michelsberger
Kultur im Bayerischen Donautal und die Michelsberger Kultur im ostlichen
Alpenvorland. Bayerischen Vorgeschichtsblätter 57, 1-31.
Martin Socas, D., Camalich Massieu, M.D. & Gonzalez Quintero, P. 1998.
L’Andalousie. In J Guilaine (ed.) Atlas du Néolithique 2B. Europe Occidentale, pp
871-934. Liège: University of Lèige
McDonnell, J.G. 2001. Pyrotechnology. In D. Brothwell and M. Pollard (eds.)
Handbook of Archaeological Sciences. pp 493-506. London: Wiley.
Meredith, Craig, 1998 An Archaeometallurgical Survey of Ancient Tin Mines and
Smelting sites in Spain and Portugal. British Archaeological Reports, International
Series 714. Oxford: Archaeopress.
Meurkens, L. 2004. A Matter of Elites, Specialists and Ritual? Social and Symbolic
dimensions of metalworking in the North-west European Bronze Age. M.A. Thesis,
University of Leiden.
Mille B. and Bouquet L., 2004. Le métal au 3e millénaire avant notre ère dans le
Centre-Nord de la France. In Vander Linden M. and Salanova L. (eds). Le troisième
millénaire dans le nord de la France et en Belgique. Actes de la journée d’études
SRBAP – SPF. 8 mars 2003, Lille. Mémoire de la Société Préhistorique Française,
XXXV, Anthropologica et Praehistorica, 115 : 197-215.
Molina, F., Cámara, J.A., Capel, J., Nájera, T & Sáez, L. 2004. Los Millares y la
periodización de la prehistoria reciente del sureste, pp 142-148. Simposos de
Prehistoria Cueva de Nerja. Nerja: Fundacion Cueva de Nerja.
Monge Soares, A. Araujo, M de F., and Peixoto Cabral, J. 1994. Vestígos da Práctica
de Metalurgia em povoados Calcolíticos da Bacia do Guadiana, entre Ardila e o
Chança. In J. Campos, J. Perez and F. Gomez (eds) Arqueologia en el entorno del
Bajo Guadiana, pp 165-200. University de Huelva: Grupo de Investigacion
Arqueológica del Património del Suroeste
Montero Ruiz, I. 1994. El origen de la Metalurgia en el Sureste Peninsula: Instituto
36
de Estudios Almerienses.
Montero Ruiz, I., Rihuete, C. & Ruiz Taboada, A. 1999. Precisones sobre el
enterramiento colectivo neolítico del Cerro Virtud (Cuevas de Almanzora, Almería).
Trabajos de Prehistoria 56(1) , 119-130.
Montero Ruiz, I. 2005. Métallurgie ancienne dans la Péninsule Ibérique. In P.Ambert
and J. Vaquer (eds) La première métallurgie en France et dans les pays limitrophes,
pp 187-194. Mémoire de la Société Préhistorique Française 37.
Much, M. 1893. Die Kupferzeit in Europa. Zena
Muhly, J., 2002 Early Metallurgy in Greece and Cyprus. In U. Yalçin (ed.) Anatolian
Metal II, pp 77-82. Der Anschnitt, Beiheft 15. Bochum: Deutches Bergbau- Museum.
Müller, Johannes, 2002 Modelle zur Einführung der Zinnbronzetechnologie und zur
sozialen Differenzierung der mitteleuropäischen Frühbronzezeit. In Vom
Endneolithikum zur Frühbronzezeit: Muster sozialen Wandels? Johannes. Müller ed
pp 267-289. Universitätsforschung zur prähistorischen Archäologie 90. Bonn: Habelt.
Müller, R., Goldenberg, G., Bartleheim, M., Kunst, M. and Pernicka, E. 2007
Zambujal and the beginnings of metallurgy in southern Portugal. In S. La Niece, D.
Hook and P. Craddock (eds.) Metals and Mines: studies in archaeometallurgy, pp1526. London: Archetype.
Müller, R., Rehren, T. & Rovira, S. 2004. Almizaraque and the early metallurgy of
southeast Spain: new data. Madrider Mitteilungen 45: 33-56.
Nadler. M & Zeeb, A. 1994. Südbayern zwischen Linearbandkeramik und Altheim:
ein neuer Gliederungsvorschlag. In H.J. Beier (ed.) Der Rössner horizont in Mittel
Europa, pp 127-90. Wilkau-Hasslau: Beier & Beran.
Needham, S., 1996 Chronology and periodisation in the British Bronze Age. Acta
Archaeologica 67, 121-40.
Needham S., 1998. Modelling the flow of metal in the Bronze Age. In C Mordant, M
Pernot and V Rychnes (eds). L'Atelier du Bronzier en Europe du xxe au viiie siècle
avant notre ère; III, Production, Circulation et Consommation du Bronze. pp. 285307. Paris: CTHS.
Needham, S., 2001. When expediency broaches ritual intention: the flow of metal
between systemic and buried domains. Royal Anthropological Institute 7, 275-298.
Needham, S., 2002. Analytical implications for Beaker metallurgy in North-West
Europe. In E. Pernicka and M. Bartelheim (ed.) The Beginnings of Metallurgy in the
Old World, pp 99-133. Rahden: Verlag Marie Leidorf.
Needham, S., 2004. Migdale-Marnoch: sunburst of Scottish metallurgy. In Ian
Shepherd and Gordon Barclay (eds.) The Neolithic and Early Bronze Age of Scotland
37
in their European context, pp 217-245. Edinburgh: Royal Society of Antiquaries of
Scotland.
Nocete, F. 2001. Tercer Milenio Antes de Nuestra Era: relaciones y contradiciones
centre/periferia en el Valle del Guadalquivir. Barcelona: Bellaterra Arqueología.
Nocete, F. (ed.) 2004.Odiel: Projecto de Investigación Arqueológia para el análisis
de la desigualidad social en el Suroeste de la Península Ibérica. Seville: Monografias
de Arqueologia 19.
Nocete, F. 2006. The first specialised copper industry in the Iberian peninsula:
Cabezo Juré (2900-2200 BC). Antiquity 80, 646-657.
Northover, Peter 1989. Properties and Use of Arsenic-Copper Alloys. In Ernst
Pernicka and Gerhard Wagner (eds.) Old World Archaeometallurgy, Andreas
Hauptmann,. pp 111-118. Der Anschnitt Beiheft 7. Bochum: Deutsches Bergbau
Museum.
Northover, P. O’Brien, W. & Stos, S. 2001. Lead isotopes and metal circulation in
Beaker/Early Bronze Age Ireland. Journal of Irish Archaeology 10, 25-48.
O’Brien, W. 2004. Ross Island (Bronze Age Studies 5). Galway: National University
of Ireland.
Orton, C., Tyers, P. and Vince, A. 1997. Pottery in Archaeology. Cambridge:
Cambridge University Press.
Ottaway, B., 1973. The earliest copper ornaments in northern Europe.
Proceedings of the Prehistoric Society 39: 294-331.
Ottaway, B.S. 1994. Prähistorische Archäometallurgie. Verlag Marie L. Leidorf,
Espelkamp.
Ottaway, B.S. 2001. Innovation, production and specialisation in early prehistoric
copper metallurgy. European Journal of Archaeology 4 (1), 87-112.
Ottaway, B.S. 2003. Experimental Archaeometallurgy. In T. Stöllner, G. Steffens & J.
Cierny (eds) Man and Mining - Mensch and Bergbau: studies in honour of Gerd
Weisgerber on occasion of his 65th birthday, pp 341-48. Der Anschnitt Beiheft 9.
Bochum: Bergbau-Museum.
Ottaway, B.S. & Roberts, B. 2008. The Emergence of Metalworking. In A. Jones
Prehistoric Europe. London: Blackwell.
Ottaway, B. & S. Seibel 1998. Dust in the Wind: experimental casting of
bronze in sand moulds. In M-C Frère-Sautot (ed) Paléométallurgie des cuivres: Actes
du colloque de Bourg-en-Bresse et Beaune, 17-18 oct. 1997, pp 59-63. Monographies
Intrumentum 5. Montagnac: Monique Mergoil
Pare, C., 2000 Bronze and the Bronze Age. In C. Pare, (ed.) Metals Make the World
38
go Round, pp 1-38. Oxford: Oxbow Books.
Pascale De, A., 2003 “Hammerstones from early copper mines” sintesi dei
ritrovamenti nell’ Europa e mel Mediterraneo orientale e prime considerazioni sui
mazzuoli di Monte Loreto (IV millennio BC – Liguria). Rivista di Studi Liguri 69, 542.
Pellicer, M. & Hurtado, V. 1980. El Poblado Metalúrgico de Chinflón (Zalamacea la
Real, Huelva). Seville: Publicaciones del Departmento de Prehistoria y Arqueología.
Perea, A. 1991. Orfebrería Prerromana: Arqueologia de Oro. Madrid: Caja de
Madrid
Pingel, V. 1992. Die vorgeschichtlichen Goldfunde der Iberischen Halbinsel:
Eine archäeologische Untersuchung zur Auswertung der Spektranalysen. Deutsches
Archäologisches Institut. Madrider Forschungen, Band 17. Berlin: Walter de Gruyter
und Co.
Pernicka, E., Begemann, F., Schmitt-Strecker, S., & Wagner, G.A. (1993) Eneolithic
and Early Bronze Age copper artifacts from the Balkans and their relation to Serbian#
copper ores. Praehistorische Zeitschrift 68: 1-54.
Pernicka, E. 1998. Die Ausbreitung der Zinnbronze im 3. Jahrtausend. In Mensch und
Umwelt in der Bronzezeit Europas, Bernd Hänsel ed. pp 135-147. Kiel: Oetker-Voges
Verlag.
Pétrequin, P., 1993. North wind, south wind. Neolithic technical choices in the Jura
mountains, 3700-2400 BC. In P. Lemonnier (ed.), Technological choices.
Transformation in material cultures since the Neolithic, pp 36-76. London, Routledge.
Pigeot, N. 1990 Technical and social actors. Flintknapping specialists and apprentices
at Magdalenian Etiolles. Archaeological Review from Cambridge 9 (1): 126-141.
Prange, M., Ambert, P. & C. Strahm 2003 Geochemical and lead isotope
characterisation of copper ores and metal objects from Cabrières/Hérault, Languedoc,
S. France. In Archaeometallurgy in Europe pp 283-292. Milan: Associazione Italiana
di Metallurgia
Prange, M. & Ambert, P. 2005. Caractérisation géochimique et isotopique des
minerais et des métaux base cuivre de Cabrières (Hérault). In P.Ambert & J. Vaquer
(eds) La première métallurgie en France et dans les pays limitrophes, pp 71-81.
Mémoire de la Société Préhistorique Française 37.
Primas, M., 1995. Gold and Silver during the 3rd mill cal BC. In G. Morteani and J.P.
Northover (eds) Prehistoric Gold in Europe, pp 77-93. Dortrecht: NATO ASI Series.
Primas, M., 2002. Early Tin Bronze in Central and Southern Europe. In M.
Bartelheim, E. Pernicka & R. Krause The Beginnings of Metallurgy in the Old World.
eds. pp 303-314. Rahden/Westf: Marie Leidorf.
39
Radivojevic, M. 2007. Evidence for early copper smelting in Belovode, a Vinča
culture site in eastern Serbia. London: UCL MSc Thesis.
Rehren in press From Mine to Microbe: the Neolithic copper melting crucibles from
Switzerland. In T Rehren, I Freestone and A Shortland (eds) Festschrift for Michael
Tite, pp 223-233. UCL: UCL University Press.
Renfrew, C. 1967. Colonialism and Megalithismus. Antiquity 41, 276-288.
Renfrew, C. 1973. Before civilization: the radiocarbon revolution and prehistoric
Europe.Cambridge: Cambridge University Press.
Roberts, B. in press a Migration, Craft Expertise and Metallurgy: analysing the
‘spread’ of metal in Europe. Archaeological Review from Cambridge
Roberts, B. in press b Creating Traditions and Shaping Technologies: understanding
the earliest metal objects and metal production in Western Europe. World
Archaeology
Roden, C 1988 Blasrohrdüsen: Ein archäologischer Exkurs zur Pyrotechnologie des
Chalkolithikums und der Bronzezeit. Der Anschnitt 40: 62–82.
Rohl, B. & Needham, S. 1998. The Circulation of Metal in the British Bronze Age:
the application of lead isotope analysis. British Museum Occasional Paper 102.
Rothenberg, B. & Andrews P. 1996. Prehistoric Copper Mining in South-West Spain:
the evidence from Chinflón. Unpublished Report
Rothenberg B. & A. Blanco-Freijeiro 1980. Ancient Copper mining and smelting at
Chinflon (Huelva, SW Spain). London: British Museum Occasional Paper 20, 41-62.
Rothenberg, B. & A. Blanco-Freijeiro 1981. Ancient Mining and Metallurgy in
South-west Spain. London: I.A.M.S.
Rovira, S. 1998. Metalurgia campaniforme en España: resultados de quince años de
investigación arqueometalúrgia. In Marie-Chantal Frère-Sautot (ed.) Paléométallurgie
des cuivres: Actes du colloque de Bourg-en-Bresse et Beaune, 17-18 oct. 1997, pp
109-127. Monographies Intrumentum 5. Montagnac: Monique Mergoil
Rovira, S., 2002. Metallurgy and society in Prehistoric Spain. In B.S. Ottaway and
E.C. Wager (eds.) Metals and Society: papers from a session held at the European
Association of Archaeologists Sixth Annual Meeting in Lisbon 2000, pp 5-20. British
Archaeological Reports (International Series) 1061. Oxford: Archaeopress.
Rovira, S and Delibes de Castro, G. 2005. Beaker metallurgical technology in the
Iberian Peninsula: pouring, moulding and post-smelting treatments. In M. A Rojo
Guerra, R. Garrido Pena, I. García-Martínez de Lagrán (ed.) El Campaniforme en la
Península Ibérica y Su Contexto Europeo, pp 513-21. Valladolid: Universidad de
Valladolid.
40
Rovira, S. and Guttierez, A. 2005. Utilisation expérimentale d’un four primitif pour
fonder du minerai de cuivre. In P.Ambert and J. Vaquer (eds) La première
métallurgie\en France et dans les pays limitrophes, pp 241-246. Mémoire de la
Société Préhistorique Française 37.
Roussot-Larroque, J. 2005. Premiere metallurgie du Sud-Ouest Atlantique de la
France. In P.Ambert & J. Vaquer (eds) La première métallurgie en France et dans les
pays limitrophes, pp 159-174. Mémoire de la Société Préhistorique Française 37.
Rovira, S. & P. Ambert, 2002a. Les céramiques à réduire minerales de cobre: une
technique métallurgique utilisée en Ibérie, son extension en France méridonale.
Bulletin de la Société Préhistorique Française 99 (1), 105-126.
Rovira, S. & Ambert, P. 2002b. Vasijas cerámicas para reducir minerales de cobre en
a Península Ibérica y en la Francia Meridional. Trabajos de Prehistoria 59 (1), 89105.
Rovira, S. & P. Gómez-Ramos. 2004. Las Primeras Etapas Metalurgicas en la
Peninsula Ibérica. III. Estudios Metalográficos. Madrid: Instituto Universitario
Ortega y Gasset y Ministerio de Educacion, Cultura y Deporte.
Rovira, S. & Guttierez, A. 2005. Utilisation expérimentale d’un four primitif pour
fonder du minerai de cuivre. In P.Ambert & J. Vaquer (eds) La première métallurgie
en France et dans les pays limitrophes, pp 241-246. Mémoire de la Société
Préhistorique Française 37.
Rowlands, M. 1971. The archaeological interpretation of metalworking. World
Archaeology 3, 210-223.
Ruíz Taboada, A. and I. Montero-Ruíz. 1999. The oldest metallurgy in western
Europe. Antiquity 73, 897-903.
Ryndina, N., Indenbaum, G. and Kolosova, V. 1999. Copper Production
from polymetallic sulphide ores in the northeastern Balkan Eneolithic Culture.
Journal of Archaeological Science 26: 1059-1068
Sáez, R. Nocete, F., Nieto, J.M., Ángeles Capitán, M. and Rovira, S. 2003. The
extractive metallurgy of copper from Cabezo Juré, Huelva: chemical and
mineraological study of slags dated to the third millennium BC. The Canadian
Mineralogist 41(3), 627-638
Shennan, S. 1993: Commodities, transactions and. growth in the Central-European
early Bronze Age. Journal of European Archaeology 1, 59–72
Shennan, S. 1999. Cost, benefit and value in the organization of early European
copper production. Antiquity 73, 352-63.
Sherratt, A. 1976. Resources, technology and trade: an essay on early European
metallurgy. In G.Sieveking, I.Longworth, and K.Wilson Problems in Economic and
Social Anthropology, pp. 557-581. London: Duckworth
41
Sherratt, A. 1994. What would a Bronze Age world system look like? Relations
between temperate Europe and the Mediterranean in later prehistory. Journal of
European Archaeology 1, 1-57.
Siret, E. & L. Siret. 1887. Les Premiers Ages du Metal dans le sud-est de l’ Espagne.
Antwerp
Skeates, Robin, 1994. Early metal use in the Central Mediterranean region. Accordia
Research Papers 4: 5-47.
Sofaer, J. 2006. Pots, Houses and Metal: Technological Relations at the Bronze Age
Tell at Százhalombatta, Hungary. Oxford Journal of Archaeology 25 (2) 127-147.
Sofaer-Derevenski, J. and M.L. Stig-Sørensen. 2002. Becoming Cultural: Society and
the Incorporation of Bronze. In B.S. Ottaway and E.C. Wager (eds) Metals and
Society: papers from a session held at the European Association of Archaeologists
Sixth Annual Meeting in Lisbon 2000, pp 117-122. British Archaeological Reports
(International Series) 1061. Oxford: Archaeopress.
Sofaer-Derevenski, J. and M.L. Stig-Sørensen in press Technological change as social
change: the introduction of metal in Europe. In M. Bartelheim and V. Heyd (eds)
Continuity-discontinuity: transition periods in European prehistory. Forschungen zur
Archäometrie und Altertumswissenschaft 3. Rahden:
Srejovic, D. 1972. Lepenski Vir. London: Thames & Hudson.
Strahm, C. 1994. Die Anfänge der Metallurgie in Mitteleuropa. Helvetia
Archaeologica 25, 2-39.
Strahm, C. 2005. L’introduction et la diffusion de la métallurgie en France. In P.
Ambert & J. Vaquer (eds) La première métallurgie en France et dans les pays
limitrophes, pp 27-36. Mémoire de la Société Préhistorique Française 37.
Suarez, A., Mellado, C., Ortiz, D. & San Martín, C. 1986. Aportaciones al estudio de
la Edad del Bronce en la provincia de Almería: análisis de la distribución de
yacimientos. In Homenaje a Luis Siret, pp 196-207. Seville: Junta de Andalucia.
Taylor, T., 1999. Envaluing metal: theorizing the Eneolithic 'hiatus'. In S. Young, M.
Pollard, P. Budd and R. Ixer (eds.) Metals in Antiquity, pp 22-32 British
Archaeological Reports (International series) 792. Oxford: Archaeopress.
Thornton, C.P. 2002. The Domestication of Metal: a reassessment of the early use of
copper minerals and metal in Anatolia and southeastern Europe. M.Phil. Thesis.
Cambridge University.
Timberlake, S. 2002. Ancient prospection for metals and modern prospection for
ancient mines: the evidence for Bronze Age Mining within the British Isles. In M.
Bartelheim, E. Pernicka and R. Krause (ed.) The Beginnings of Metallurgy in the Old
World, pp 327-357. Rahden: Verlag Marie Leidorf.
42
Timberlake, Simon 2003. Excavations on Copa Hill, Cwmystwyth (1986-1999); An
Early Bronze Age copper mine within the uplands of Central Wales. British
Archaeological Reports (British Series) 348. Oxford: Archaeopress.
Timberlake, S. 2005. In search of the first melting pot. British Archaeology 82,
32-33.
Timberlake, S. 2007. The use of experimental archaeology/archaeometallurgy for
understanding and reconstructing Early Bronze Age mining and smelting. In Susan La
Niece, Duncan Hook and Paul Craddock eds. Metals and Mines: studies in
archaeometallurgy pp 27-36. London: Archetype and British Museum
Tylecote, R. 1970. The composition of metal artifacts: a guide to provenance?
Antiquity 44,19-25.
Tylecote, R. 1987. An Early History of Metallurgy in Europe. London: Longman.
Vander Linden, M. 2006. Le phénomène campaniforme dans l'Europe du 3ème
millénaire avant notre ère: Synthèse et nouvelles perspectives. Oxford: British
Archaeological Reports (International Series) 1470.
Vander Linden, M. 2007. What linked Bell Beakers in third millennium BC Europe?
Antiquity 81, 343-52.
Wailes, B., (ed.) 1996. Craft Specialization and Social Evolution: in memory of V.
Gordon Childe. University Museum Monograph 93, Philadelphia: University
Museum of Archaeology and Anthropology, University of Pennsylvania.
Warmenbol E., 2004. Le début des âges des Métaux en Belgique. In Vander
Linden M. & Salanova L. (eds) Le troisième millénaire dans le nord de la France et en
Belgique. Actes de la journée d’études SRBAP – SPF. 8 mars 2003, Lille. Mémoire
de la Société Préhistorique Française, XXXV, Anthropologica et Praehistorica, 115 :
27-48.
Waterbolk, H.T. & J.J. Butler. 1965. Comments on the use of metallurgical analysis in
prehistoric studies. Helenium 5, 227-241.
Wayman, M. & J. Duke 1999 The Effects of Melting on Native Copper. In A.
Hauptmann, E. Pernicka,T. Rehren & Ü. Yalçin (eds.). The Beginnings of Metallurgy,
pp 55-63. Bochum: Der Anschnitt.
Weisgerber, G. & Pernicka, E. 1995. Ore mining in prehistoric Europe: an overview.
In G. Morteani & J. P. Northover, (eds.) Prehistoric Gold in Europe, pp. 159-182.
Netherlands: Kluwer Academic Publishers.
Weisgerber, G. & Willies, L. 2001. The use of fire in prehistoric and ancient mining:
firesetting. Paleorient 26(2), 131-149.
Wilk, R. 2001. Toward at archaeology of needs. In M. Schiffer, (ed.) Anthropological
43
Perspectives on Technology, pp 107-122. Albuquerque: University of New Mexico
Press.
Zachos, K. 2007. The Neolithic background: a reassessment. In P.M. Day and R.C.P.
Doonan (eds.) Metallurgy in the Early Bronze Age Aegean, pp 168-206. Sheffield:
Sheffield Studies in Aegean Archaeology 7.
Figures
Figure 1a
Major Copper sources in Europe
44
Figure 1b
Major gold sources in Europe
45
Figure 1c
Lead ores and argentiferous lead ores in Europe
46
Figure 1d
Major Tin ore sources in Europe
Figure 2
Clay ringed hearth structure at Los Millares, southeast Spain
(Craddock 1995, 133)
47
Figure 3
Typical early copper objects found in southern Iberia featuring
daggers (48-52), a flat axe (53), a chisel (55) and an awl (54)
(Martin Socas et al. 1998, 923)
48