Anda di halaman 1dari 9

Journal of Archaeological Science 36 (2009) 308316

Contents lists available at ScienceDirect

Journal of Archaeological Science


journal homepage: http://www.elsevier.com/locate/jas

The production of speiss (iron arsenide) during the Early Bronze Age in Iran
C.P. Thornton, Th. Rehren*, V.C. Pigott
UCL Institute of Archaeology, 3134 Gordon Square, London WC1H 0PY, United Kingdom

a r t i c l e i n f o
Article history: Received 18 June 2008 Received in revised form 27 August 2008 Accepted 5 September 2008 Keywords: Metallurgy Early Bronze Age Speiss Iron Arsenic Tepe Hissar Iran

a b s t r a c t
In this paper, analyses of some unusual slag samples from the prehistoric site of Tepe Hissar in northeastern Iran are presented. These slags are the remains of a ve-thousand-year-old pyrotechnological process that produced speiss, a quasi-metallic material usually formed as an accidental by-product of copper or lead smelting. We argue that the speiss slags from Tepe Hissar suggest the intentional production of ironarsenic alloy (speiss) in prehistory. Why the Tepe Hissar metalworkers produced speiss is a question that requires further investigation, but our preliminary assessment suggests that it was to provide arsenic as an alloying component for arsenical copper, the preferred copper alloy during much of the Early Bronze Age in Iran, and widely used across the ancient world. This recognition signicantly advances our understanding of the early stages of metallurgy in the Old World. 2008 Elsevier Ltd. All rights reserved.

1. Introduction Many archaeometallurgical studies are focused exclusively on metal artefacts which, through chemical, isotopic and microscopic analyses, can reveal ancient alloying, casting and working practices, and often the ore sources utilized. However, these nished products represent only a small fraction of the total metallurgical assemblage. Indeed, a typical prehistoric metallurgical site contains a combination of technical nds such as stone tools (e.g., hammer stones), ore and gangue pieces, fragments of technical ceramics (e.g., crucibles, furnaces, moulds, tuyeres), charcoal, and various types of slag. By targeting these waste products, we can potentially learn a great deal more about production and distribution than by simply targeting the consumed nal products (Bachmann, 1982; see also papers in Shimada, 2007). This paper focuses on a particular and not generally recognised type of waste material called speiss slag. Our use of the term speiss slag is modelled on the term copper slag, which is the siliceous waste material discarded after smelting copper ore. Our use of speiss slag stands therefore in contrast to other uses in which speiss slag is an intermediary stage in ancient metallurgical processes (e.g., Paulin et al., 2000, who refer to the speiss itself with this term).

2. What is speiss? There are two main types of speiss to be considered in archaeometallurgy. Ferrous speiss is typically a mixture of arsenical iron and iron arsenides, while base-metal speiss is typically a complex mixture of copper, nickel, iron and/or silver as arsenides and antimonides, often with some sulphur and lead. On a materials level, speiss refers to intermetallic compounds consisting of transition metals and elements of the fth main group of the Periodic Table of Elementsdi.e., N, P, As, Sb and Bi. In practice, only arsenic and antimony appear regularly in metallurgical speiss, with phosphorous playing a role as alloying component in early iron, and very occasionally in lead smelting based on phosphate minerals. Both arsenic and antimony were used as alloying elements for copper, and there is a continuum between arsenical/antimonial copper (with all arsenic and antimony in solid solution in the copper alpha phase), arsenical and antimonial bronze (comprising the alpha solid solution together with some intermetallic compounds), and speiss proper (containing little or no solid solution alpha phase). Lead, which is not a transition metal, does not form intermetallic compounds with arsenic or antimony, and as such is not formally part of speiss; however, due to its geochemical association with many base metal ores and its low melting point, it often accompanies the speiss phase when it forms. Base-metal speiss is known from a number of prehistoric and historic metal smelting sites, forming especially when smelting complex arsenic- or antimonyrich copper or lead ores (see Bachmann, 1982: 2930; Keesmann, 1999; Rehren et al., 1999). The formation of ferrous speiss is not as well understood, as most ferrous speiss objects have been found

* Corresponding author. Tel.: 44 20 7679 4757; fax: 44 20 7383 2572. E-mail address: th.rehren@ucl.ac.uk (Th. Rehren). 0305-4403/$ see front matter 2008 Elsevier Ltd. All rights reserved. doi:10.1016/j.jas.2008.09.017

C.P. Thornton et al. / Journal of Archaeological Science 36 (2009) 308316

309

outside their production context, thus making the ferrous speiss slags presented here so signicant (Fig. 1). Speiss is a material similar in appearance to metal, although it is brittle and cannot be worked other than by casting. Thus, it is usually discarded as a waste product. Given that a signicant amount of copper can be trapped in the speiss (see Lechtman and Klein, 1999: 508509), the occurrence of speiss increases metal loss during copper smelting. In the case of lead smelting as a preliminary step in silver production, the occurrence of speiss is considered particularly detrimental as much of the silver will get trapped in the speiss instead of in the lead metal, whence it may be separated via cupellation (Craddock et al., 1987). In central and western Europe, copper- and lead-rich speiss was produced as early as the Chalcolithic period (Keesmann and Mor ller et al., 2004). Occasionally, base-metal eno-Onorato, 1999; Mu speiss was traded in ingot form during the Late Bronze Age (Paulin et al., 1999, 2000; Paulin and Orel, 2003) and Roman period (Tylecote, 1987: 199201; Craddock et al., 1987, contra Kassianidou, 1998), possibly to be used for the further extraction of copper, lead, and/or silver trapped therein. In the medieval period, speiss is generally understood to have been an unwanted material (Goldenberg, 1996; Rehren et al., 1999), although the discovery recently of large quantities of copper- and nickel-rich speiss ingots from a late medieval ship wreck in the Baltic (D. Noreus, pers. commun. 2008) may indicate that it was still considered useful enough to be traded. From the ancient Near East to Central Asia, published evidence for even the accidental production of speissdlet alone the use of it dis limited. However, it should be noted that many elite metal artefacts of the 4th and 3rd millennia in the Levant (e.g., Nahal Mishmar), in Anatolia (e.g., Arslantepe VIA), and in Trans-Caucasia (e.g., Shengavit) contain up to 25 wt% arsenic and/or antimony (Tadmor et al., 1995; Palmieri et al., 1999; Meliksetian et al., 2003). Under equilibrium conditions, both copper-antimony and copperarsenic form pure intermetallic phases with around 30 wt% Sb and As (and the formulae Cu3Sb and Cu2As, respectively). Thus, these alloys with more than 20 wt% arsenic and/or antimony could be

Fig. 1. In an ideal copper smelt involving complex sulpharsenide ores, the metalworker achieves near-perfect segregation between four major phases, according to their respective specic gravities: (1) silicate slag, (2) copper-rich matte (i.e., copper (iron) sulphide), (3) speiss, and (4) copper metal. Simplied from Keesmann (1991).

dened as bordering on base-metal speiss. In addition, the production of a copper-arsenic-lead alloy with around 15 to 20 wt% arsenic during the rst millennium BC has been reported by Mei (2000: 5657) from Nulasai in Xinjiang, western China, together with the analysis of a copper-arsenic-lead buckle found in the Hami region, containing some 25 wt% arsenic (Mei, 2000: 40). The melting temperature of these alloys is comparatively low, with the binary eutectic for copper-arsenic and copper-antimony both melting at less than 700  C. The presence of signicant amounts of lead will further lower the melting temperature and will increase the freezing range for alloys near the eutectic. These properties likely made speiss-rich alloys very suitable for complex castings, such as those from the Nahal Mishmar hoard (e.g., Shalev and Northover, 1993). The formation of ferrous speiss, dened here as iron arsenide with <23 wt% base-metal content, obviously involves the smelting of ores rich in iron and arsenic. However, we know little about the economic, cultural, geological or metallurgical context within which this would have taken place. Usual explanations of the accidental smelting of arsenopyrite make little sense for societies that had been smelting specic ores successfully for centuries. There are a number of reported occurrences of ferrous speiss, the earliest dating from the 3rd millennium BCE at Shahr-i Sokhta, eastern Iran (Hauptmann et al., 2003). Doonan et al. (2007: 109) report a piece of partly corroded plate-like iron arsenide found at Early Bronze Age Poros-Katsambas in northern Crete, with 53 wt% Fe and 39 wt% As, intergrown with another phase showing the typical eutectic speiss texture known from Shahr-i Sokhta. Other examples of early ferrous speiss include a lump of heattreated ironarsenic material with some sulphur from a copper smelting workshop in EB/MB Jericho (Khalil and Bachmann, 1981), ironarsenic ingots from iron-working areas at mid-2nd millenazko y in central Turkey (Muhly et al., 1985), a similar nium Bog fragment from the Late Bronze Age copper and lead workshop at Kamid el-Loz in Lebanon (Frisch et al., 1985: 146148), and another from a LH IIIB workshop at Tiryns in Greece (Kilian, 1983: 304; Waldbaum, 1999: 31). More recently, ferrous speiss was identied in a Graeco-Roman context in Kythera (Broodbank et al., 2007). Some of these examples were found among ordinary iron metal stock rather than among waste material such as slag. This may indicate that ferrous speiss was seen originally as another kind of ferrous metal, suitable for casting but not for smithing, which might explain Hittite texts describing gurines made of iron (Muhly et al., 1985: 73). What is interesting about all of these ferrous speiss examples, and the ferrous speiss slags presented herein, is that none appears to contain signicant amounts of base-metals, which one would expect to nd if they derived from copper or lead smelting. Of the examples given above for which analyses have been published, the late 2nd millennium speiss from Kamid el-Loz has the greatest amount of these metals, which point analysis suggests is probably no more than w2 wt% of the total (Frisch et al., 1985: Tafel 76). In contrast, the iron-rich speiss found in early 3rd millennium slag from Almizaraque in southeast Spain, which probably derived from ller smelting copper arsenate ores, has more like 20 wt% Cu (Mu et al., 2004: 44). Such low levels of copper and lead suggest that early ferrous speiss was produced outside base metal smelting, possibly in a failed attempt to make iron using arsenic-rich iron ores, and then either used for casting or simply discarded. However, this interpretation of speiss as a failed product of iron smelting does not make sense in an Early Bronze Age context. Instead, the speiss was perhaps produced by smelting iron arsenide ores such as llingite (FeAs2), or blue-green scorodite arsenopyrite (FeAsS), lo (FeAsO4-2H2O) and then either used for casting as a metal in its own right or, as will be discussed further below, to be used as an alloying component with copper to produce arsenical copper.

310

C.P. Thornton et al. / Journal of Archaeological Science 36 (2009) 308316

Doonan et al. (2007: 109, 111113) interpret their EBA nd from llingite mineral used to increase the amount of Crete as a piece of lo arsenic in the local copper in a crucible-bound alloying step. Their archaeological context and the evidence for both increased iron content in archaeologically-related crucible slags and increased arsenic content in copper prills is convincing for this claim, although we interpret the primary iron arsenide phase in their nd llingite should contain about 27 wt% Fe differently. The mineral lo and 73 wt% As; their analysis, with 53 wt% Fe and 39 wt% As, is more consistent with the intermetallic compound Fe2As (theoretically 60 wt% Fe and 40 wt% As). This ironarsenide phase is not known in mineralogy but is a typical speiss phase with a melting point of around 930  Cdwell within the capabilities of ancient metalworkers. Thus, it is likely that this iron arsenide represents speiss, a metallurgical product, rather than a mineral. Its origin is unknown at present, but there seem to be no arsenic mineralizations known on Crete available to ancient metalworkers. It is against this background of limited understanding of the role of speiss in early metallurgy that the nds presented in this paper are highly signicant. 3. The Tepe Hissar speiss slags The speiss slags presented here derive from the 1976 restudy project at the site of Tepe Hissar near the modern city of Damghan in northeastern Iran (see Dyson and Howard, 1989) (Fig. 2). Tepe

Hissar is a multi-period site beginning in the late 5th millennium BCE with seemingly continuous occupation until the early 2nd millennium BCE. In the 4th and perhaps again in the 3rd millennium BCE, Hissar was a major processing site for precious stones such as lapis lazuli and alabaster, which were then sent further west to Mesopotamia and beyond (Tosi, 1974; Majidzadeh, 1982). In addition, the site is covered with metallurgical remains including slag, prills, crucible fragments, and furnace lining (Pigott et al., 1982; Pigott, 1989), suggesting it joined contemporary sites on the Iranian Plateau such as Arisman (Chegini et al., 2000, 2004; Helwing, 2005), Sialk (Nezafati and Pernicka, 2006), Ghabristan (Majidzadeh, 1979, 1989), and later Shahr-i Sokhta (Hauptmann and Weisgerber, 1980; Artioli et al., 2005) in producing large quantities of copper, copper alloys, lead, and silver for both local consumption and long-distance trade. Recently, eighteen examples of slag from 4th-early 3rd millennium contexts on the Main Mound of Tepe Hissar have been studied; of these, three are speiss slags (Table 1). One of these was found in two parts (H76-S30 and -S39); their chemical and textural similarity as well as their shared archaeological context led us to treat them as a single piece although the two fragments do not join together. The three speiss slags were excavated during the 1976 season near to a mud-brick platform on which slag, crucible fragments, furnace lining, prills, moulds, iron oxide minerals, and other debris were found. The speiss slag fragments were found in two different stratigraphic phases dated by the excavators to ca. 33503100 BCE

Fig. 2. Map of Iran showing some major sites mentioned in the text.

C.P. Thornton et al. / Journal of Archaeological Science 36 (2009) 308316 Table 1 Bulk chemical composition of the speiss slags and slags from the same contexts Sample S120 MS161a MS161b MS161c S63a S63b S86 S46 S48 S30 S39 S43 S40b S40a S47a S47b S47c S47d S41a S41b Date (ca.) 3600 BCE Method SEM XRF SEM SEM XRF SEM XRF XRF SEM XRF XRF XRF XRF SEM XRF XRF SEM SEM XRF SEM Na2O 0.00 1.46 0.92 0.31 1.36 0.24 1.15 2.15 1.37 0.81 1.22 1.04 1.88 1.40 1.09 2.51 2.32 2.19 0.32 2.65 MgO 0.76 3.30 3.26 2.88 1.46 2.23 1.67 1.95 2.73 0.88 1.23 1.64 1.02 2.96 1.08 2.13 0.94 2.68 1.54 2.79 Al2O3 5.25 9.03 13.38 12.40 7.68 7.91 8.59 9.91 11.85 8.13 5.56 11.22 5.30 11.15 2.04 7.13 4.61 10.84 6.96 6.98 SiO2 37.9 28.7 56.1 53.3 34.7 51.2 35.5 37.6 59.6 34.0 28.7 55.8 28.2 54.8 25.2 31.7 50.7 52.7 54.0 43.7 P2O5 0.00 0.14 0.00 0.00 0.13 0.00 0.37 0.30 0.00 0.24 0.31 0.15 0.90 0.27 0.54 0.24 0.00 0.12 0.09 0.00 SO3 0.27 1.37 0.05 0.14 0.35 0.73 0.00 0.00 0.12 2.67 1.85 0.13 5.59 0.26 6.39 0.12 0.04 0.35 0.03 3.82 K2O 0.52 1.23 2.82 2.64 1.11 2.06 1.18 1.64 2.76 0.82 0.83 1.66 0.97 3.02 0.79 0.96 0.35 2.01 0.98 0.63 CaO 2.15 10.05 16.3 12.55 4.45 10.96 14.67 12.91 15.61 7.33 4.08 7.57 7.11 16.39 10.53 10.20 7.56 14.14 3.26 5.44 TiO2 0.05 0.38 0.84 0.66 0.40 0.54 0.32 0.43 0.77 0.31 0.33 0.38 0.25 0.79 0.25 0.33 0.24 0.70 0.21 0.32 MnO 0.00 1.27 0.00 0.27 1.94 0.71 0.98 0.81 0.00 0.07 0.06 0.15 0.05 0.00 0.04 0.67 0.31 0.07 0.29 0.00 FeO 50.0 37.0 5.5 7.9 39.1 16.0 32.4 29.3 4.8 42.3 54.8 14.2 47.5 5.0 49.0 36.5 28.6 12.6 24.7 13.0 CuO 2.68 2.58 0.61 6.41 2.27 6.47 0.48 0.79 0.13 0.06 0.01 2.39 0.02 2.83 0.02 3.96 2.93 1.06 3.40 11.98 ZnO 0.00 0.07 0.00 0.00 0.11 0.00 0.21 0.09 0.00 1.49 0.16 2.41 0.25 0.00 0.28 0.34 0.00 0.00 1.24 0.15 As2O3 0.00 0.04 0.00 0.00 0.00 0.00 0.02 0.00 0.00 0.13 0.43 0.02 0.17 0.38 1.91 0.15 0.00 0.00 0.06 0.19 SrO 0.00 0.15 0.00 0.00 0.20 0.00 0.12 0.24 0.00 0.16 0.11 0.07 0.39 0.00 0.32 0.19 0.00 0.00 0.11 0.00 BaO 0.00 3.03 0.00 0.00 4.22 0.00 2.06 1.45 0.00 0.33 0.15 0.55 0.17 0.00 0.14 0.59 0.00 0.00 0.20 6.23 PbO 0.00 0.05 0.00 0.00 0.16 0.00 0.04 0.06 0.00 0.08 0.01 0.20 0.01 0.49 0.09 1.33 0.00 0.00 2.48 0.00 Cl wt% 0.31 0.10 0.13 0.41 0.29 0.89 0.12 0.20 0.22 0.04 0.05 0.26 0.04 0.14 0.09 0.86 1.25 0.42 0.07 2.03

311

FeO:SiO2 1.32 1.29 0.10 0.15 1.13 0.31 0.91 0.78 0.08 1.24 1.91 0.26 1.68 0.09 1.94 1.15 0.57 0.24 0.46 0.30

3400 BCE

3200 BCE

3000 BCE

3000 BCE

2800 BCE

All data normalized to 100 wt%. Bold indicates a speiss slag.

(H76-S30/S39) and ca. 31002900 BCE (H76-S40b and H76-S47a). It is possible that, given their proximity to each other, these slag pieces are actually from one chronological period (although the later two were found above a clearly-dened oor located roughly half a meter higher than the surface of the mud-brick platform). However, we would argue based on their chemical and mineralogical composition that they are different enough to suggest at least two if not three distinct smelting events, even if they were actually contemporary. It should also be remembered that the 1976 excavations were limited in scope and extentdhad successive excavation seasons been permitted, perhaps more samples would have been uncovered. Most slag samples were analysed as pressed pellets for bulk chemical composition using a Spectro Xlab 2000 ED-XRF. Pieces too small for XRF analysis were polished and analysed using a Hitachi S-3400N scanning electron microscope with energy-dispersive spectrometer. In all cases, a sample from each slag was mounted in resin and analysed for phase identication on a Leica DMLM optical microscope. All analyses were carried out at the Wolfson Archaeological Science Laboratories at the UCL Institute of Archaeology. (1) H76-S30/S39 (CG90 5 lot 13, ca. 33503100 BCE) is made up of two pieces of tap slag that were found atop the metalworking platform mentioned above. The two fragments are black, covered in rust-brown iron hydroxides, heavy, dense (as opposed to porous), and strongly magnetic. Microscopic analysis of S30 shows a fairly homogenous iron-rich slag with a quenched exterior surface, a quickly-cooled and more oxidized area just inside the quenched surface, and a slower-cooled and more reduced interior section (Fig. 3). S39 appears to be a continuation of this interior part of the smelt, supporting the assertion that these two objects are from a single slag. The major silicates are skeletal lathes of fayalite (Fe2SiO4 with small amounts of Mg, Ca, and Zn), although iron-rich pyroxenes (i.e. (Fe, Mg)CaSi2O6) are found in the more quenched areas. Speiss prills (with 5070 wt% As, 1838 wt% Fe, 39 wt% Ni, 37 wt% Cu, and small amounts of Co) in the more oxidized areas are low in sulphur (15 wt% S), although there are still a few small iron sulphide prills (with w18 wt% Cu and w2.5 wt% Zn). In the more reduced areas of the interior, the prills are small and exclusively composed of iron sulphides (sometimes with minor amounts

of As and Cu). A few partially-reacted fragments of gangue were found in S39, including quartz and iron oxide. A different slag fragment discarded on the same metalworking platform (H76-S43) is also a tap slag, but it is not magnetic, displays numerous vacuoles, and is covered in various copper oxides. Analysis of this slag, which has many large and small prills of copper (with <8.3 wt% As and <5.1 wt% Fe) and matte (with <1.2 wt% Fe and small inclusions of As-Cu-Ni-Sb-Co-sulphide), shows high levels of silica (see Table 1) and even relict pieces of quartz gangue. It seems highly unlikely that this slag and the speiss slags came from the same smelt, although they may originate from smelting related ores, perhaps from the same complex ore source. Their presence in the same archaeological context underlines the varied character of the metallurgy at Tepe Hissar. (2) H76-S40b (CG90 P lot 13, ca. 31002900 BCE) is dark grey, massive with a few large vacuoles, heavy, magnetic, and displays no visible oxides on the surface. The microstructure of the slag is fairly homogenous and consists principally of fayalite crystals with small amounts of Mg and Ca (Fig. 4). S40b is more corroded than S30/S39; thus, all of the larger prills have been entirely oxidized. The tiny prills in the glassy matrix surrounding the fayalite crystals are predominantly pure iron sulphides, while the slightly larger iron sulphide prills contain up to 5 wt% As. S40b was found in association with a crucible rim sherd (H76S40a) displaying a glassy slag with a few small prills of copper (with <2.7 wt% Fe and <5.8 wt% As) and copper matte (with <2 wt% As and <1 wt% Fe). This crucible appears to have been used for the production of arsenical copper. The fact that iron does not appear to have been added to the slag beyond what came from the ceramic (see Table 1) suggests either the mixed smelting of copper oxide and arsenic sulphide ores (e.g., realgar (AsS) or orpiment (As2S3)) or perhaps the rening of arsenic-rich copper matte with low levels of iron. In general, the low levels of iron in the slag preclude it from being related to either the production or the use of ferrous speiss. (3) H76-S47a (CG90 P lot 12, ca. 31002900 BCE) is grey, massive with a few large vacuoles, heavy and magnetic, and covered in secondary iron hydroxides. The sample is quite corroded and was not as internally homogenous as the other three. The slag is

312

C.P. Thornton et al. / Journal of Archaeological Science 36 (2009) 308316

Fig. 4. Photomicrograph of well-formed fayalite crystals and round Fe(As)-sulphide prills in H76-S40b. Width of image 1.5 mm.

slag sample. The prills in S47a are mostly ironarsenic-sulphides, containing w4550 wt% Fe, 3550 wt% As, and 520 wt% S. In one part of the slag sample (Fig. 6), a partially-reacted piece of quartz gangue was found with inclusions of iron sulphide (with 12 wt% As) and lead sulphide (with <12 wt% Fe and <2.6 wt% As) located amongst various Fe-spinels and lathes of fayalite. This is probably indicative of the type of complex ore bodies from which the arsenic originated, and suggests the use of (iron) arsenic sulphide ores in a (self-uxing) quartz gangue for the production of speiss. S47a was found in association with three other metallurgical nds: a copperleadarsenic smelting slag (H76-S47b), a solitary example of copper plattenschlacke (H76-S47c), and a crucible fragment for copper smelting or rening (H76-S47d). Of these three, only S47b could be in some way related to S47a by its arsenic content; however, the presence of large prills of copper (with w4 wt% As and 3 wt% Pb), lead chlorides/oxides (with <15 wt% Cu, <15 wt% As, and <25 wt% Fe), and copper (iron) matte in S47b set it sufciently apart from S47a to be from a different metallurgical process. The slag itself is internally heterogeneous and is comprised

Fig. 3. Back-scattered electron image of the three zones in H76-S30. (a) Quenched surface with iron corrosion products (far right) on top of an iron-oxide skin (white). Dendrites of fayalite (light grey), Fe-rich pyroxenes (dark grey), and iron oxides (white). (b) Quickly-cooled intermediate zone with fayalitic dendrites. The bright prill is (Ba, Sr)-sulphate and the crack is lled with iron corrosion products. (c) Slowercooled interior zone with skeletal lathes of fayalite and round prills of Fe(As)-sulphide.

composed of various iron-rich spinels and well-formed crystals of fayalite and kirschsteinite (i.e., FeCaSiO4) with small amounts of Mg (Fig. 5). A fragment of the furnace or crucible lining was trapped in the slag and upon analysis revealed higher amounts of lime and calcium phosphate (i.e., apatite) than in other refractory ceramics from the site, thereby explaining the high levels of calcium in this

Fig. 5. Back-scattered electron image of well-formed (Fe,Ca)2SiO4 crystals (dark grey) with iron spinels (light grey) often overlying them in H76-S47a. The light grey feature at the bottom is a corroded speiss prill.

C.P. Thornton et al. / Journal of Archaeological Science 36 (2009) 308316

313

Fig. 6. Photomicrograph of the partially-reacted quartz gangue with adhering Pbsulphides and Fe(As)-sulphides in H76-S47a. Width of image = 3 mm.

of fayalite needles amongst Fe-rich pyroxenes and clusters of Feoxides (with Al, Mg, and Ti). The fragment of plattenschlacke (H76S47c) is also of interest in that it suggests an entirely new method of producing relatively pure copperdeither through processing of matte or direct smelting of pure (rather than mixed) copper ores dat the beginning of the 3rd millennium BCE. This production of pure copper instead of complex copper alloys, which is also supported by the crucible fragment (S47d) with its prills of pure copper, parallels the situation at Arslantepe VIB.2 (ca. 29002750 BCE) in Eastern Anatolia. Here, slags suggest the production of pure copper, although nished artefacts inexplicably contain 1 6 wt% arsenic (Palmieri et al., 1999; Hauptmann and Palmieri, 2000). We will return to this point below. 4. Interpretation The analytical data separate the slags into three main groups (Table 1). The crucible slags are particularly low in iron oxide

(less than about 15 wt%) and have high alumina values of typically more than 10 wt%, often exceeding their iron oxide content. This is consistent with their origin predominantly as fused crucible material. Their iron oxide to silica ratio is consistently below 0.4 (Fig. 7). They are contaminated with mechanically trapped copper, in the extreme reaching more than 10 wt% copper oxide. The copper slags proper have intermediate iron oxide to silica ratios (mostly 0.51.3), between 25 and 50 wt% iron oxide, less than 10 wt% alumina, and from 0.5 to 4 wt% copper oxide. Finally, the speiss slags have virtually no copper oxide (on average only about 200 to 300 ppm), from 40 to more than 50 wt% iron oxide, and always below 10 wt% alumina; their iron oxide to silica ratio ranges from 1.2 to nearly 2.0. The chemical and mineralogical analyses of the three speiss slags from Tepe Hissar presented above provide us with several general points of interest. First, all of the speiss slags are iron-rich slags. This is unusual in the context of the larger corpus of slags from the Main Mound, in that most of the other analysed copper slags from late-4th millennium contexts are predominantly silicarich (see Table 1). Second, although two of the speiss slags (S30/S39 and S47a) have similarities to copper-base slags found in the same context (e.g., S43, which has high As and Fe in copper prills and speiss inclusions in the matte), all of the speiss slags are very low in copper and lead, with a maximum of 0.06 wt% copper oxide in the bulk composition. The copper slags, in contrast, have on average more than 2 wt% copper oxide (see Table 1). Thus, we nd it difcult to imagine a scenario in which a single smelting process could produce two different classes of slag (i.e., ferrous speiss slag vs. copper slag). Perhaps future experimentation will provide a scenario in which this can happen, but for now we must assume that the speiss slags and the copper-base slags did not originate from the same smelting process. Thus, any similarities between the two may be the result of using different ores from the same general ore bodies, with some overlapping ore mineralogy, but smelting them separately. Third, the speiss slags were formed in a reducing atmosphere and under high temperatures that were maintained for an extended period of time before the smelt was allowed to cool down slowly (S40b and S47a) or tapped (S39/S30). Such conditions would

Main Mound Slags


2 1.8
S40b S47a S39

Copper slags Speiss slags Crucible slags

1.6 1.4

FeO : SiO2

1.2 1

S30

S120 MS161a S63a S47b

S86

0.8 0.6 0.4 0.2 0 0.00

S46 S47c S41a S43 S40a S63b MS161c S41b

S47d S48 MS161b

2.00

4.00

6.00

8.00

10.00

12.00

CuO
Fig. 7. Copper oxide versus FeO:SiO2 ratio, separating the three main slag groups from Tepe Hissar. Crucible slags have an iron oxide to silica ratio of less than 0.4 and can reach excessively high copper oxide levels; copper slags have typically intermediate ratios of 0.5 to 1.3 and between 0.5 and 4 wt% copper oxide, while the speiss slags have virtually no copper and very high iron oxide to silica ratios of 1.21.9. See Table 1 for numerical data.

314

C.P. Thornton et al. / Journal of Archaeological Science 36 (2009) 308316

allow for the near-complete separation of the (quasi)metallic speiss and the slag, especially if the latter was tapped out before adding more ore to the furnace to increase the nal yield. Both the reducing conditions and the well-controlled temperature would be essential for a high arsenic yield in the resulting speiss, and could explain the low levels of arsenic in the silicate matrix of all of these slags. It is no doubt signicant that the end of the 4th millennium is the time in which furnaces began to replace crucibles for smelting processes in northern Iran (see Thornton, in press). Thus, we expect that speiss slags were produced entirely in furnaces and not in crucibles, although more research is needed on this subject before rm statements can be made. Finally, it should be noted that the majority of the prills in all three slags are predominantly iron arsenic sulphides or iron sulphides, and not speiss per se. This mirrors the widespread observation from copper slags, which often contain numerous copper sulphide prills but rarely copper metal prills. In both cases, this is probably due to the fact that slags are preferentially enriched with sulphidic rather than metallic phases as a result of the latters higher specic gravity. In speiss slags, sulphides remain as a consequence of having to use reducing conditions to keep the arsenic from oxidizing. As demonstrated by the high-As and lowFe/S prills in the quenched area of S30, sulphur in speiss will oxidize before arsenic. Thus, arsenic losses are likely limited as long as sufcient sulphur remains in the system. At the end of the process, most of the sulphur will form a separate matte phase, of lower specic gravity than the speiss, and should separate easily from it (Fig. 1). The presence of pure iron sulphide prills in these slags is mirrored by the regular presence of iron sulphide inclusions in those ferrous speiss pieces that have been sufciently analysed to identify their presence or absence (e.g. those from Shar-i Soktha or Kythera; see above); thus, the presence of iron sulphide seems a common feature of ferrous speiss. This consistency not only strengthens our argument for an intentional and regular production of speiss, and relates the speiss slag reported here to typical ferrous speiss, but also gives some indication of the nature of the ore. The absence of copper beyond mere traces makes it highly unlikely that the speiss slag formed during smelting base metal ores. More likely, the ore was dominated by arsenopyrite, possibly intergrown with some pyrite. Associations of arsenopyrite with pyrite are the most common arsenic mineralizations, and arsenicrich parts could be selected by their whiter colour than the more llingite or scorodite would yellow pyrite-rich parts. Smelting pure lo be less likely to result in the formation of iron sulphide inclusions, llingite is rare as a pure since both are nominally free of sulphur. Lo mineral, but often occurs intergrown with arsenopyrite and as such may have been part of the ore charge. Both corrode supercially to scorodite, the green to blue colour of which may have rst made this assemblage attractive to ancient metallurgists. 5. Discussion Based upon our preliminary assessment of the recently-analysed slags from Tepe Hissar, it is certain that from the earliest metalworking phase attested at the site (ca. 3600 BCE), copper-base alloys were being produced through mixed- or co-smelting operations utilizing complex oxidic and sulphidic ores. If we assume that the heirs to this metalworking tradition were the same people producing speiss (as suggested by the context), then it is not difcult to imagine the metalworker smelting partly-corroded arsenopyrite-rich ore, possibly intergrown with iron hydroxide and iron arsenates (e.g, scorodite), in a self-uxing quartz gangue in a furnace. At strongly reducing and sufciently hot conditions, this combination should produce arsenic-rich speiss at the bottom; arsenic-containing matte in the middle; and iron-rich/arsenic-poor

fayalitic speiss slag on top. The relative proportions of each phase will depend on the starting composition of the ore charge and the furnace conditions. If ferrous speiss was the intended or major product, one would not expect to nd much of it among the slag fragments. If it was a minor by-product, the same would probably apply, making it difcult to decide about intentionality based only on the archaeological record. If, however, the speiss was an unintentional by-product, then one has to ask what the main product of the smelting operation would have been, given the nature of the waste material and the archaeological context of the nds. We argue that the low base metal content of these slags rules out their connection to copper or lead smelting, while their early date excludes a relationship to bloomery iron smelting. Having ruled out the traditional metals, it is reasonable to assume that the ferrous speiss was indeed the intended product of this operation. However, it is important to consider that at suitable conditions, one could even expect to produce arsenical iron metal below the arsenic-rich speiss (Muhly et al., 1985: 77; Keesmann, 1991, 1999), a point which we will return to below. The technicalities which led to the formation of this speiss slag are relatively easily reconstructed; why, however, would one want to produce ferrous speiss in the rst place? At present, no nished objects made of this alloy are known; instead, it may have been used as an alloying component in the production of arsenical copper. For this, the arsenic-rich speiss could be melted with copper (or copper matte) in a crucible under oxidizing conditions, removing iron through slagging. Although some arsenic would be lost in the process as As2O3 smoke, enough would combine with the copper to produce arsenical copper, as demonstrated by Zwicker (1991: 333334). Alternatively, the metalworker could have utilized the low boiling point of arsenic (614  C) to transfer the arsenic from the speiss to the surface of a copper-base object in a closed crucible under reducing conditions through a process known as cementation. Such silvering practices using arsenic are attested from 3rd millennium metalwork, such as the bull standard from Horoztepe (Smith, 1973). This schema for the production of arsenical copper through combining ferrous speiss and copper metal must remain speculative until further experimental and archaeological research has been carried out. However, the evidence for 3rd millennium copper production at Shahr-i Sokhta in eastern Iran shows an unusual discrepancy between the high levels of arsenic in many of the nished copper-base artefacts versus the near-complete absence of arsenic in most of the slags or copper ores from the site (see Heskel, 1982; Artioli et al., 2005). It is an open question as to how arsenic was added to the copper, but the presence of arsenic-rich ferrous speiss at this site (Hauptmann et al., 2003) is perhaps a strong clue. Further away, the evidence for alloying copper with ferrous speiss in EBA Crete provides even more technical details about how this may have been achieved (Doonan et al., 2007: 111113). Another explanation is more speculative. Both speiss and arsenical iron are relatively easily smelted and may have been used as a castable metal. No such objects are known at present from the Early Bronze Age, which may simply reect their limited corrosion resistance. The occurrence of ferrous speiss among later hoards of iron objects, listed earlier in this paper, could point to the occasional use of iron arsenic alloys in antiquity, and hence possibly in the Early Bronze Age. Clearly, further research both in the eld and in the lab are necessary to address this issue. 6. Conclusion The occurrence of speiss slags presented here, in the context of a site that had been smelting complex ores for hundreds of years, may suggest the intentional production of arsenic-rich ferrous speiss. Whether the speiss was used as an alloying agent or as

C.P. Thornton et al. / Journal of Archaeological Science 36 (2009) 308316

315

a castable material in its own right is unknown, but the evidence from Shahr-i Sokhta (Hauptmann et al., 2003) and Poros (Doonan et al., 2007) may suggest that the impetus for producing ferrous speiss on the Iranian Plateau came from a desire for a more reliable source of arsenic than the usual green copper ores which may or may not contain arsenic. It has already been shown how arsenical copper was the alloy of choice on the Iranian Plateau well into the 2nd millennium BCE (Stech and Pigott, 1986; Thornton et al., 2002), and the rise of large, lowland 3rd millennium cities like Shahr-i Sokhta and Shahdad in eastern Iran undoubtedly led to an increased demand for arsenical copper. This demand could have been fullled by importing ingots of ferrous speiss from smaller highland sites located closer to the arsenic-bearing ores, such as Tepe Hissar. Although rm evidence for the large-scale production of speiss at sites such as Hissar is still missing, only 45 of the thousands of slag fragments from the site have been analysed, and only from two small areas of the 12 ha site. There are few known sites on the Iranian Plateau comparable to Tepe Hissar in terms of the amount of production relative to size and location in the highlands. Thus, it remains to be seen whether Hissar is merely one of a number of small sites producing complex metal alloys (including speiss) in the Late Chalcolithic and Early Bronze Age. Preliminary analysis by Ernst Pernicka of some brown slags from the slag heaps of Arisman A (ca. 31002900 BCE), located near the modern city of Kashan (w300 km southwest of Hissar), has discovered iron-rich siliceous slags containing prills of iron arsenides in magnetite with little copper content (Pernicka, pers. commun. 2007). In his opinion, roughly 10% of the estimated 2030 tons of slag in the Arisman A slag heaps are brown slags. If future analysis proves this estimation to be correct, then we would be looking at a large-scale, regular and certainly intentional production of arsenical speiss in north-central Iran at the very llner (2004, Note 3) tentatively beginning of the Bronze Age. Sto suggested the use of speiss as an additive to copper smelting, based on the nds from Shar-i Sohkta and Arisman. Interestingly, the situation in Anatolia is quite similar to that in Iran, with highland sites such as Horoztepe, Norsuntepe, Arslantepe, and others demonstrating large amounts of metalworking debris relative to their small size in the late 4th-early 3rd millennium BCE (Yakar, 2002; Bilgi, 2004). As mentioned above, there is a shift to the production of relatively pure copper in the early 3rd millennium BCE (contemporary with the spread of Kura-Araxes or Early Trans-Caucasian cultures from the Caucasus) at sites such as Arslantepe, although the nished artefacts continue to include signicant levels of arsenic, nickel, and other alloying elements (see Yener, 2000: 4857). For this reason, the presence of slag containing prills of nickel-arsenic speiss from Arslantepe as reported by Hess (in Palmieri et al., 1999: 145) is of considerable interest, as it may suggest shared technological practice between these two distant regions during the Early Bronze Age. As a nal point, the most striking and perhaps controversial conclusion to be drawn from the speiss slags of Tepe Hissar is that here is the earliest evidence for the intentional smelting of iron minerals. Although the desired product was probably arsenic and not iron, it is entirely reasonable to assume that arsenical iron, or an iron-like alloy, could have been produced accidentally in the process. It seems unlikely that such an unintended product would have been discarded in ignorance, given the ample evidence for both meteoritic and terrestrial iron in the Near East before the 2nd millennium BCE (Waldbaum, 1999). Thus, if ferrous speiss production became widespread and formalized in the 3rd millennium in order to provide arsenical copper, Iran may also have been one of the heartlands of iron metallurgy. This is supported by the reported presence of late 3rd millennium cast iron from Geoy tepe (Burton-Brown, 1950) and contemporary Mesopotamian textual references to iron coming from Anshan (i.e., Tal-e Malyan, near

modern-day Shiraz in southern Iran) (Moorey, 1999: 287). Indeed, the current excavator of Malyan informs us that iron rings were found in lot 27 of sounding H1s in Kaftari contexts dating to the very beginning of the 2nd millennium BCE (John Alden, pers. commun. 2007). Much has yet to be done before these hypotheses can be accepted as conclusions, but the growing body of evidence suggests that ferrous speiss was being produced on the Iranian Plateau (and, perhaps, in Anatolia) by the early 3rd millennium BCE. The intended purpose and the long-term effects of this technological practice remain to be seen, and only future archaeological and archaeometallurgical research will prove or disprove what we have proposed here about the production of arsenical copper and the origins of iron metallurgy. Until then, the speiss slags from Tepe Hissar demonstrate effectively how much new information about the development of metallurgy can be gleaned from analysing waste products and not just nished artefacts. Acknowledgements We are indebted to Kevin Reeves, Simon Groom and Sandra Bond at the UCL Institute of Archaeology for technical support. C.P.T. acknowledges nancial support from the European Union, Marie Curie EST Action Science and Conservation in Archaeology, Contract MEST-2004-519504 for a three-months fellowship at the UCL Institute of Archaeology in 2006. Thanks to David Killick and numerous anonymous reviewers for their comments and criticisms which helped to improve the manuscript. Bibliography
Artioli, G., Giardino, C., Guida, G., Lazzari, A., Vidale, M., 2005. On the exploitation of copper ores at Shahr-i Sokhta (Sistan, Iran) in the 3rd millennium BC. In: Franke-Vogt, U., Weisshaar, H.-J. (Eds.), South Asian Archaeology 2003. For ologie Auereuropa ischer Kulturen (FAAK), Band 1. Verlag schungen zur Archa Linden Soft, Bonn, pp. 179184. Bachmann, H.-G., 1982. The Identication of Slags from Archaeological Sites. Institute of Archaeology, London. Occasional Publication No. 6. ., 2004. Anatolia, Cradle of Castings. Do ktas, Istanbul. Bilgi, O Broodbank, C., Rehren, Th., Zianni, A., 2007. Scientic analysis of metal objects and metallurgical remains from Kastri, Kythera. Annuals of the British School in Athens 102, 221240. Burton-Brown, T., 1950. Iron objects from Azarbaijan. Man 50, 79. llner, T., Chegini, N.N., Momenzadeh, M., Parzinger, H., Pernicka, E., Sto Vatandoust, A., Weisgerber, G., 2000. Preliminary report on archaeometallurgical investigations around the prehistoric site of Arisman near Kashan, western Central Iran. Archaeologische Mitteilungen aus Iran und Turan 32, 281318. historische Chegini, N.N., Helwing, B., Parzinger, H., Vatandoust, A., 2004. Eine pra Industriesiedlung auf dem iranischen Plateau-Forschungen in Arisman. In: llner, T., Slotta, R., Vatandoust, A. (Eds.), Persiens Antike Pracht. Deutsches Sto Bergbau-Museum, Bochum, pp. 210216. Craddock, P.T., Freestone, I.C., Hunt Ortiz, M., 1987. Recovery of silver from speiss at Rio Tinto (SW Spain). IAMS Newsletter 10/11, 811. Doonan, R., Day, P., Dimpoulou-Rethemiotaki, N., 2007. Lame excuses for emerging complexity in Early Bronze Age Crete: the metallurgical nds from Poros Katsambas and their context. In: Day, P., Doonan, R. (Eds.), Metallurgy in the Early Bronze Age Aegean. Oxbow Books, Oxford, pp. 98122. Dyson Jr., R.H., Howard, S.M. (Eds.), 1989. Tappeh Hesar: Reports of the Restudy Project, 1976. Case Editrice le Lettere, Firenze. tten der Frisch, B., Mansfeld, G., Thiele, W.-R., 1985. Kamid el-Loz 6. Die Werksta tbronzezeitlichen Pala ste. Dr. Rudolf Habelt GMBH, Bonn. spa ometallurgische Untersuchungen zur Entwicklung des Goldenberg, G., 1996. Archa ttenwesens im Schwarzwald. Blei-, Silber- und Kupfergewinnung von Metallhu hgeschichte bis zum 19. Jahrhundert. In: Goldenberg, G., Otto, J., der Fru ometallurgische Untersuchungen zum Metall-hu ttenSteuer, H. (Eds.), Archa wesen im Schwarzwald. Thorbecke Verlag, Sigmaringen, pp. 9275. Hauptmann, A., Rehren, Th., Schmitt-Strecker, S., 2003. Early Bronze Age copper llner, T., Koerlin, G., metallurgy at Shahr-i Sokhta (Iran), reconsidered. In: Sto Steffens, G., Cierny, J. (Eds.), Man and MiningdMensch und Bergbau. Studies in Honour of Gerd Weisgerber: Der Anschnitt, Beiheft 16. Deutsches BergbauMuseum, Bochum, pp. 197213. Hauptmann, A., Palmieri, A.M., 2000. Metal production in the Eastern Mediterranean at the Transition of the 4th/3rd millennium: Case studies from Arslantepe. . (Ed.), Anatolian Metal I. Der Anschnitt, Beiheft 13. Deutsches In: Yalcin, U Bergbau-Museum, Bochum, pp. 7582.

316

C.P. Thornton et al. / Journal of Archaeological Science 36 (2009) 308316 Metallurgy. Der Anschnitt, Beiheft 9. Deutsches Bergbau-Museum, Bochum, pp. 141149. Paulin, A., Orel, N.T., 2003. Metallurgical examinations in the archaeometallurgical projects of the National Museum of Slovenia. Materiali in Tehnologije 37 (5), 251259. Paulin, A., Spaic, S., Spruk, S., Heath, D.J., Trampuz-Orel, N., 1999. Speiss from Late Bronze Age. Erzmetall 52 (11), 615622. Paulin, A., Spaic, S., Heath, D.J., Trampuz-Orel, N., 2000. Analysis of Late Bronze Age Speiss. Bulletin of the Metals Museum 32, 2941. Pigott, V.C., 1989. Archaeo-metallurgical investigations at Bronze Age Tappeh Hesar, 1976. In: Dyson, R.H., Howard, S.M. (Eds.), Tappeh Hesar: Reports of the Restudy Project, 1976. Case Editrice le Lettere, Firenze, pp. 2534. Pigott, V.C., Howard, S.M., Epstein, S.M., 1982. Pyrotechnology and culture change at Bronze Age Tepe Hissar (Iran). In: Wertime, T.A., Wertime, S.F. (Eds.), Early Pyrotechnology: The Evolution of the First Fire-Using Industries. Smithsonian Institution Press, Washington DC, pp. 215236. Rehren, Th., Schneider, J., Bartels, Ch., 1999. Medieval lead-silver smelting in the Siegerland, West Germany. Historical Metallurgy 33, 7384. Shalev, S., Northover, P., 1993. The metallurgy of the Nahal Mishmar hoard reconsidered. Archaeometry 35, 3547. Shimada, I. (Ed.), 2007. Craft Production in Complex Societies: Multicraft and Producer Perspectives. University of Utah Press, Salt Lake City. Smith, C.S., 1973. An examination of the arsenic-rich coating on a bronze bull from Horoztepe. In: Young, W.J. (Ed.), Application of Science in the Examination of Works of Art. Boston Museum of Fine Arts, Boston, pp. 96102. Stech, T., Pigott, V.C., 1986. The metals trade in southwest Asia in the third millennium B.C. Iraq 48, 3964. llner, T., 2004. Prehistoric and ancient ore-mining in Iran. In: Sto llner, T., Sto Slotta, R., Vatandoust, A. (Eds.). Persiens Antike Pracht. Deutsches BergbauMuseum, Bochum, pp. 4463. Tadmor, M., Kedem, D., Begemann, F., Hauptmann, A., Pernicka, E., SchmittStrecker, S., 1995. The Nahal Mishmar hoard from the Judean Desert: technology, composition, and provenance. Atiqot 27, 95148. Thornton, C.P., in press. The emergence of complex metallurgy on the Iranian Plateau: Escaping the Levantine Paradigm. Journal of World Prehistory. Thornton, C.P., Lamberg-Karlovsky, C.C., Liezers, M., Young, S.M.M., 2002. On pins and needles: tracing the evolution of copper-base alloying at Tepe Yahya, Iran, via ICP-MS analysis of common-place items. Journal of Archaeological Science 29, 14511460. Tosi, M., 1974. The lapis lazuli trade across the Iranian Plateau in the 3rd millennium B.C. In: Gururajamanjarika: Studi in onore di Guiseppe Tucci. Instituto Universitario Orientale, Napoli, pp. 322. Tylecote, R.F., 1987. The Early History of Metallurgy in Europe. Longman, New York. Waldbaum, J.C., 1999. The coming of iron in the eastern Mediterranean: Thirty years of archaeological and technological research. In: Pigott, V.C. (Ed.), The Archaeometallurgy of the Asian Old World. University Museum Monograph 89. The University Museum, University of Pennsylvania, Philadelphia, pp. 2757. Yakar, J., 2002. East Anatolian metallurgy in the fourth and third millennia BC: Some remarks. In: Yalcin, U. (Ed.), Anatolian Metal II. Der Anschnitt, Beiheft 15. Deutsches Bergbau-Museum, Bochum, pp. 1526. Yener, K.A., 2000. The Domestication of Metals. Brill, Leiden. Zwicker, U., 1991. Natural copper-arsenic alloy and smelted arsenic bronzes in early metal production. In: Mohen, J.-P., Eluere, C. (Eds.), Decouverte du Metal. Picard, Paris, pp. 331340.

Hauptmann, A., Weisgerber, G., 1980. The Early Bronze Age copper metallurgy of orient 6, 120127. Shahr-i Sokhta (Iran). Pale Helwing, B., 2005. Early mining and metallurgy on the western Iranian Plateau: First results of the Iranian-German archaeological research at Arisman, 2000 2004. Archaeologische Mitteilung aus Iran und Turan 37, 425434. Heskel, D.L., 1982. The development of pyrotechnology in Iran during the fourth and third millennia B.C. PhD Dissertation, Department of Anthropology, Harvard University. Kassianidou, V., 1998. Was silver actually recovered from speiss in antiquity? In: Rehren, Th., Hauptmann, A., Muhly, J.D. (Eds.), Metallurgica Antiqua. Der Anschnitt, Beiheft 8. Deutsches Bergbau-Museum, Bochum, pp. 6976. Khalil, L., Bachmann, H.-G., 1981. Evidence of copper smelting in Bronze Age Jericho. Historical Metallurgy 15, 103106. Keesmann, I., 1991. Rio Tinto: Die Technik der Silbergewinnung zu Beginn des Mittelalters, in: Conference papers, Archeologie, Sous theme Argent, plomb et cuivre dans lhistoire. CNRS, Villeurbanne, pp. 113. dwestiberischen SulphiderzKeesmann, I., 1999. Eisen in antiken Schlacken des su rtels. Pallas 50, 339360. Gu Keesmann, I., Moreno-Onorato, A., 1999. Naturwissenschaftliche Untersuchungen hen Technologie von Kupfer und Kupfer-Arsen-Bronze. In: Hauptmann, A., zur fru . (Eds.), The Beginnings of Metallurgy. Pernicka, E., Rehren, Th., Yalcin, U Der Anschnitt, Beiheft 9. Deutsches Bergbau-Museum, Bochum, pp. 317332. Kilian, K., 1983. Ausgrabungen in Tiryns 1981. Bericht zu den Grabungen. Arch ologischer Anzeiger 3, 294328. a Lechtman, H., Klein, S., 1999. The production of copper-arsenic alloys (arsenicbronze) by cosmelting: modern experiment, ancient practice. Journal of Archaeological Science 26, 497526. Majidzadeh, Y., 1979. An early coppersmith workshop at Tepe Ghabristan, in: Akten r Iranische Kunst und Archa ologie, des VII. Internationalen Kongresses fu nchen 710 September 1976. Archaeologische Mitteilungen aus Iran, Berlin, Mu suppl. 6, pp. 8292. orient 8 (1), Majidzadeh, Y., 1982. Lapis lazuli and the Great Khorasan Road. Pale 5969. Majidzadeh, Y., 1989. An early industrial proto-urban center on the central plateau of Iran: Tepe Ghabristan. In: Leonard, A.J., Williams, B.B. (Eds.), Essays in Ancient Civilization Presented to Helene J. Kantor. Studies in Ancient Oriental Civilization, 47, Chicago, pp. 157-166. Mei, J., 2000. Copper and Bronze Metallurgy in Late Prehistoric Xinjiang. BAR IS 865. Meliksetian, K., Pernicka, E., Badaylan, R., Avetissyan, P., 2003. Geochemical characterization of Armenian Early Bronze Age metal artefacts and their relation to copper ores. In: Archaeometallurgy in Europe. Associazione Italiana di Metallurgia, Milano, pp. 597-606. Moorey, P.R.S., 1999. Ancient Mesopotamian Materials and Industries: The Archaeological Evidence, second ed. Eisenbrauns, Winona Lake. zgen, E., 1985. Iron in Anatolia and the nature of Muhly, J.D., Maddin, R., Stech, T., O the Hittite iron industry. Anatolian Studies 35, 6784. ller, R., Rehren, Th., Rovira, S., 2004. Almizaraque and the early copper metalMu lurgy of southeast Spain: New data. Madrider Mitteilungen 45, 3356. Nezafati, N., Pernicka, E., 2006. The smelters of Sialk: Outcomes of the rst stage of archaeometallurgical researches at Tappeh Sialk. In: Shahmirzadi, S.M. (Ed.), The Fishermen of Sialk. Archaeological Report Monograph Series, 7. Iranian Center for Archaeological Research, Tehran, pp. 79102. Palmieri, A.M., Frangipane, M., Hauptmann, A., Hess, K., 1999. Early metallurgy at Arslantepe during the Late Chalcolithic and the Early Bronze Age IA-IB periods. . (Eds.), The Beginnings of In: Hauptmann, A., Pernicka, E., Rehren, Th., Yalcin, U

Anda mungkin juga menyukai