Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Archaeological Science 101 (2019) 52–62

Contents lists available at ScienceDirect

Journal of Archaeological Science


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

Questioning Fe isotopes as a provenance tool: Insights from bog iron ores T


and alternative applications in archeometry
Thomas Rosea,∗, Philippe Téloukb, Sabine Kleinc, Horst R. Marschalla
a
Institut für Geowissenschaften, Goethe Universität, Altenhöferallee 1, 60438, Frankfurt am Main, Germany
b
Univ. Lyon, ENS-Lyon, Université Lyon, Lyon, France
c
Forschungsbereich Archäometallurgie, Deutsches Bergbau-Museum Bochum, Am Bergbaumuseum 31, 44791, Bochum, Germany

A R T I C LE I N FO A B S T R A C T

Keywords: Provenancing metal artifacts with scientific methods is an established tool in archeometry to identify the metal
Iron metallurgy deposits, which were exploited for the production of ancient metal objects. It is thus an important method to
Provenancing reconstruct ancient exchange networks and the socio-economic organization of past societies. Previously in-
Bog iron ore troduced isotope-based approaches (Pb, Sr, Os) and chemical analyses of slag inclusions have severe limitations
Iron isotopes
concerning their application or the amount of sample required. Fe isotopes were previously suggested in a quite
MC-ICP-MS
Smelting
early stage of investigation as a potential provenance tracer and it was postulated that they would not fractionate
Organic matter during the smelting procedure. However, previously published analytical data from iron ores indicate a wide
overlap between deposits. Further, the earlier studies on Fe isotopes did not included bog iron ores, despite their
high importance in ancient metallurgy. As geochemical reactions during ore formation are complex and frac-
tionation cannot be generally excluded, the applicability of Fe isotopes as an alternative provenance tracer still
asks for further investigation.
This purely methodological study focuses on specimens from two sites of the formerly mined bog iron ore
deposit Eyller Bruch (Germany), which are analyzed together with products of a smelting experiment based on
these ores. The Fe isotopic composition of the bog iron ore from the investigated region suggests an intra-deposit
zonation caused by environmental parameters and its overall variation is comparable with that of other deposits.
The bog iron ore isotope signature largely overlaps with the isotope range of mineralizations in other regions. As
a consequence of this and although the absence of Fe isotope fractionation during the smelting procedure is
confirmed, the study demonstrates the lack of discriminatory power of Fe isotopes for provenance studies.
Potential applications for archeometry can rather be found in the environmental parameters, especially organic
matter, which seems to have a strong influence on Fe isotope compositions of bog iron ores. Zonations within
deposits might be identified and could help to reconstruct the exploitation history of the deposit or to reconstruct
past bog landscapes.

1. Introduction 1.1. Current approaches to provenancing iron artifacts

The determination of the ore source of metals such as iron is a pi- Provenancing iron artifacts is still a challenging task. The loss of the
votal part in the reconstruction of the socio-economic organization of original surface by corrosion processes and limited shape variation of
ancient societies and exchange networks among them. Iron ores are tools due to functionality constraints often renders archeological ap-
more easily accessible than e. g. copper ores due to their wide-spread proaches (e. g. typology) difficult. Consequently, several scientific
occurrence, especially as surficial bog iron ore. In combination with methods for provenancing metal and metallurgical by-products were
more desirable mechanical properties, this was presumably one reason developed (Charlton, 2015), but all have some major limitations. In
for the wide spread of iron metal for tools, weapons, and other objects. contrast to other ores like for copper, lead isotopes cannot be used due
to the (often too) low concentrations and its heterogenous distribution
especially in bog iron ores (Schwab et al., 2006). Strontium isotopes
were also investigated as a potential provenance tracer, but


Corresponding author.
E-mail address: thomas.rose@daad-alumni.de (T. Rose).

https://doi.org/10.1016/j.jas.2018.11.005
Received 11 July 2018; Received in revised form 31 October 2018; Accepted 19 November 2018
Available online 30 November 2018
0305-4403/ © 2018 Elsevier Ltd. All rights reserved.
T. Rose et al. Journal of Archaeological Science 101 (2019) 52–62

compositions are largely controlled by furnace lining and not by the ore Moynier et al., 2017), the small number of studies concerning iron
(Degryse et al., 2007, 2009). At the current state osmium isotopes seem isotopes allow only a rough assessment. It seems that supergene pro-
to be a very promising tracer, but their determination requires large cesses slightly enrich the heavier isotopes in the altered products
amounts of sampled material and intricate analytical work (Brauns (Cheng et al., 2015). The interaction of primary hematite with hydro-
et al., 2013; Dillmann et al., 2017). For these reasons, the most reliable thermal fluids seems to preponderantly enrich the lighter isotopes in
and mainly used method is the chemical analysis of slag inclusions the secondary hematite. For siderite, isotopes fractionate in the oppo-
entrapped in the bloom (Blakelock et al., 2009; Coustures et al., 2003; site direction (Markl et al., 2006). Although very limited, these studies
Desaulty et al., 2009; Dillmann and L'Héritier, 2007). The main and might hint towards the possibility to (generally) distinguish among iron
trace elements of slag inclusions are measured with LA-ICP-MS deposits through their Fe isotope compositions, as pointed out by Milot
(L'Héritier et al., 2016) and multivariate statistics is used to establish a et al. (2016).
link between the slag entrapped in the bloom and the ore (e. g. Leroy
et al., 2012). However, this method necessitates experience to differ- 1.3. Fe isotope fractionation during bog iron ore formation
entiate between inclusions from smelting and forging. It also requires a
large amount of sampled material, because a sufficiently large number Especially in Northern and Middle Europe, prehistoric iron me-
of inclusions have to be analyzed (Benvenuti et al., 2016; Dillmann tallurgy widely relied exclusively on the supply of bog iron ores because
et al., 2017; Disser et al., 2014). Like for all methods relying on che- other high-quality ores were not available (e. g. Halkon, 2014; Joosten
mical or isotopic data, the determination of the ore origin relies on the et al., 1998; Larsen and Rundberget, 2014; Schwab et al., 2006;
exclusion of non-matching origins. For a successful reconstruction of Thelemann et al., 2017). Consequently, the full potential of Fe isotopes
the ore origin all ore deposits must be characterized and overlapping as a provenance tracer can only be assessed if the full isotopic variation
compositions as well as mixing renders the reconstruction of the ore of bog iron ore deposits is investigated. So far, only the processes in bog
origin difficult or even impossible. iron ore evolution were investigated but not the isotopic variability of a
Milot et al. (2016) and Milot (2016), suggested stable iron isotopes bog iron ore deposit.
as a potential tool to provenance iron artifacts. It requires only small Bog iron ore forms when reduced, Fe-rich ground water rises near
amounts of sampled material and they can routinely be analyzed. In the surface (Kaczorek and Sommer, 2003; Thelemann et al., 2017).
archeology, Fe isotopes were used only once before with the aim to Mainly through bacterial activity, Fe gets oxidized and precipitates as
identify black and red pigments (Eerkens et al., 2014). To evaluate their Fe hydroxide, which reacts to ferrihydrite (Crerar et al., 1979;
potential as a tracer in iron archeometallurgy, Milot et al. (2016) and Schwertmann et al., 1999; Teutsch et al., 2005; Yapp, 2001). During
Milot (2016) analyzed ores, metallurgical by-products and blooms from these fast reactions, the heavier isotopes are preferentially adsorbed on
two smelting experiments, as well as semi-products from the Roman the bacteria (Banning, 2008; Beard et al., 2010) and the ferrihydrite
shipwrecks of Les Saintes-Maries-de-la-Mer, France. Previous studies (Bullen et al., 2001; Frierdich et al., 2015; Mikutta et al., 2009). The
already established the latter to originate from the Montagne Noire ferrihydrite is stabilized through the adsorption of reduced Fe ions
(Baron et al., 2011; Coustures et al., 2006). They concluded that the Fe (Pedersen et al., 2005) and converts to goethite (Poage et al., 2000;
isotope composition of the ore is not altered during smelting. It was Schwertmann and Murad, 1983). This predominantly grows further
shown that the Fe isotopic composition of the semi-products agrees through adsorption of reduced Fe ions (Hansel et al., 2005). Simulta-
with ores from their suspected origin. Hence, the authors suggest a neously, preferentially light Fe isotopes from goethite are dissolved
combination of Fe isotopes with other tracing methods for the prove- (Jang et al., 2008; Poulson et al., 2005) and with time, a constant ex-
nance reconstruction of iron artifacts. change between small (< 500 nm) goethite crystals and dissolved iron
eventually balances the isotopic composition of both (Handler et al.,
1.2. Iron isotope fractionation in iron ore deposits 2009; Reddy et al., 2015; Wu et al., 2011). The latter effect is ques-
tionable for larger crystals (Handler et al., 2014), and their initial iso-
Iron has four stable isotopes, 54Fe, 56Fe, 57Fe, and 58Fe, which have topic composition might be preserved instead. Additionally, the content
natural abundances of 5.84%, 91.76%, 2.12%, and 0.28%, respectively of organic carbon (Ilina et al., 2013), plants (Fantle and DePaolo, 2004;
(Hoefs, 2009). Their distribution in terrestrial rocks is comparably Wiederhold et al., 2006), and the silicate content (Wiederhold et al.,
homogeneous (Beard et al., 2003; Poitrasson, 2006). But more surface- 2007; Wu et al., 2011) can alter the isotopic composition of goethite. As
near reduction or oxidation (redox) reactions and biological processes a result, bog iron ores show (a) a homogeneous isotope signature,
can alter the rock composition, and are capable of fractionating Fe which corresponds to some degree to the isotopic composition of the Fe
isotopes significantly (overall range of Fe isotope compositions: 3.5‰/ source or (b) they potentially possess a large variation with isotope
atomic mass unit (amu), Hoefs, 2009). Since their first analysis by patterns dependent on environmental parameters. The range of isotope
Beard and Johnson (1999), the scope of their application steadily in- composition might be even larger than those observed in hydrothermal
creased and today covers the whole range of geochemical research (see ore deposits.
the recent reviews of Albarède, 2015; Dauphas et al., 2017; Dauphas
and Rouxel, 2006). 1.4. Aim of the study
Concerning ore deposits, notable variations were observed in ar-
cheologically irrelevant banded iron formations (e. g. Frierdich et al., This entirely methodological study explores the Fe isotopic varia-
2014; Halverson et al., 2011; Steinhoefel et al., 2009) and hydro- bility of a bog iron ore deposit and aims towards a more comprehensive
thermal ore deposits (Markl et al., 2006; Wang et al., 2015; Wawryk evaluation of Fe isotopes as a provenance tracer for iron artifacts.
and Foden, 2015). Studies focusing on hydrothermal iron ore deposits Therefore it does not refer to any specific regional or timely arche-
are scarce. Before Milot et al. (2016) and Milot (2016), only Markl et al. ological context. Rather it uses own data and published data for an
(2006), Horn et al. (2006), and Cheng et al. (2015) published data for entirely scientific view on the potential of Fe isotopes as provenance
non-marine ore deposits. In such deposits isotope fractionation is tool. Although the conclusions drawn from them might seem to be
mostly driven by supergene processes. These affect surface-near rocks abstract or unrelated from an archeological point of view, they are part
and minerals along the groundwater level and are influenced by des- of the underlying framework for the application of Fe isotopes as pro-
cending atmospheric oxygen and meteoric water. Under these condi- venance tracer.
tions, the deeper and hydrothermally formed metalliferous primary Data from two different sampling regions will be directly linked to
minerals partly oxidize. While the effect of these processes on e. g. both, the Fe isotope composition of the source and the isotopic frac-
copper isotopes is well understood now (Mathur and Fantle, 2015; tionation processes during bog iron ore formation. It will be shown that

53
T. Rose et al. Journal of Archaeological Science 101 (2019) 52–62

variation of Fe isotopes within the bog iron ore deposit is comparable 2.2. Smelting experiment
with other deposit types and gives preliminary evidence for an intra-
deposit zonation. Prior to this study, M. Arians-Kronenberg (Krefeld) conducted a
Additionally, materials from a smelting experiment with bog iron successful smelting experiment with ore from the Gelderner Fleuth. She
ore from the same deposit are analyzed to verify whether Fe isotopes do generously provided specimens from the ore (ARC762), roasted ore
not fractionate during smelting. Because bog iron ores might show a (ARC761), slag (ARC760, MAK-Sl), bloom (ARC759), and a forged semi
very heterogeneous Fe isotope composition, all specimens were sam- product (MAK-Fe) for analysis. The smelting process of the ore ran
pled multiple times to investigate if the potentially strong hetero- about 6 h and a total of 30 kg (6 kg/h) were smelted. The ore:charcoal
geneous isotopic composition of the bog iron ore is preserved in the ratio was 1:1. Subsequently, 1.2 kg of the bloom was further processed
smelting products. Implications for tracing iron artifacts with Fe iso- and a piece of 700 g of purified bloom was used to forge, among others,
topes are deduced by comparing the bog iron ore data with data from the semi-product (pers. comm. M. Arians-Kronenberg, 2018).
the literature and with two Lahn-Dill type ores. Moreover, possible A polished section of the ore and the roasted ore revealed a het-
applications beyond provenancing are suggested. erogeneous structure (Fig. 2). In the ore, a lot of plant-like structures
like cross sections of stems or roots seem to be present. In a small cavity
2. Materials a ferricreted root channel was directly accessible (Fig. 2). XRD analysis
gave a mineralogical composition of siderite and goethite with more
2.1. Bog iron ores than 50% siderite. The roasted ore shows similar features and consists
of nodule-like light and dark red areas. According to XRD it consists of
Modern bog iron ores can be regarded as suitable material for the almost pure hematite with minor amounts of quartz.
study for the following reasons: (1) to rule out any influence from al-
teration processes during burial of unsmelted ore; (2) most of the pre- 2.3. Additional samples
historically exploited bog iron ore deposits are either completely
exploited or protected and accessible only with special permission; (3) Additionally to this set of material, two iron ore specimens of the
bog iron ore formation processes are mostly independent from the de- Lahn-Dill type were collected west of Wetzlar (Hesse, Germany, Fig. 1b)
posit's geographical location and had not changed over time. and analyzed. The area was extensively mined for iron ore from the 18th
Several specimens of bog iron ores were collected from two sites century until the 20thcentury. Ore-GF was gathered in the forest near
called “Gelderner Fleuth” and “Tote Rahm”. Both are part of the former the iron ore mine Grube Fortuna, today a public mine, and consists
bog iron ore mining area “Nierstal/Eyller Bruch” (Dassel, 2007; entirely of hematite. Ore-W was picked up from a small ore heap in a
Sommer, 1963) near Krefeld (North Rhine-Westphalia, Germany). The forest about 5 km southwest of Wetzlar. Fe oxides (mostly hematite)
two areas are less than 10 km apart (Fig. 1a). Ore formation began characterize this sample with a weathered surface, which is intensively
about 5 ka ago, when the Niers and its meanders were cut off from the crossed by quartz veins.
Rhine main stream. In the cut-off meanders, iron rich ground water Although initially sampled for another study, data for the metallic
from the sandy and gravel-rich lower Rhine terrace rises up to the core of the excavated semi-product SM9-99.242 (generously provided
surface, where the dissolved iron is oxidized (Dassel, 2007). by L. Long, Marseille) are also reported for comparison with the data of
This ore field was chosen because both areas provide very different Milot et al. (2016) and Milot (2016) but will not be discussed in detail
kinds of ores, which allows to assess a wide range of any intra-deposit here. It belongs to the cargo of the Roman ship wreck SM9 (14–50 AD),
variation, if present. In the nature reserve “Tote Rahm” (samples la- which is part of the famous ship wreck assemblage found in the Med-
beled ore-bog) in the Southern area of the deposit ore formation is still iterranean Sea near Les Saintes-Maries-de-la-Mer (Bouches-du-Rhône),
ongoing and the collected ore pieces consist of soft intense orange-rusty France (Long et al., 2002; Pagès, 2008; Pagès et al., 2008).
material without any visible grains (Fig. 2). The sampling location
“Gelderner Fleuth” near Nieukerk was drained and leveled in 1928, 3. Methods
which impeded further ore formation (Sommer, 1963). Two fields were
surveyed here. Although located directly next to each other and only 3.1. Sampling procedure
separated by the small river Schwarze Rahm, ore specimens from one
field are predominantly pieces consisting of sand grains with iron rich To avoid contamination from steel tools, their use was reduced to a
phases as binding material (samples labeled ore-field1, Fig. 2) while the minimum. Ore specimens were either broken in half and unweathered
specimen from the other field (labeled ore-field2) found on the river material was extracted from the interior with pieces of transparent
bank of the Schwarze Rahm consists of nodules agglomerated by iron- plastic and by rubbing the halves together (ore-bog, ore-field1), re-
rich crusts (Fig. 2). spectively, or a lancet and tweezers were used to detach small pieces
The mineralogical composition of the ores was determined with X- from the outer and interior areas of the nodules (ore-field2) and also a
ray diffraction analysis (XRD) on two nodules from each deposit (ore- small piece of a ferricreted root channel in the specimen ARC762
bog, ore-field1) and on two different areas from the specimen ore- (sample ARC762-2). Ore-bog5 consists of clay, adhering to one of the
field2. To obtain a rough semi-quantitative information on the phase ore specimen. The samples ore-bog4, 6 and 7 are from the clay filling of
compositions, Rietveld refinement was carried out with the program a cavity in a specimen, the encrusted cavity walls, and the soft material
GSAS II (Toby and Dreele, 2013). Specimens from the Tote Rahm outside of the cavity, respectively. Ore-field1-8 is surface material from
consist entirely of goethite or a mixture of goethite and lepidocrocite specimen ore-field1-7.
and less than 5% quartz. In the samples from the Gelderner Fleuth, field All other specimen were sampled with a Dremel 4000 equipped
1 more than 75% are made up by quartz or feldspar and quartz and the with a 1.2 mm diamond sputtered steel drill. To avoid contamination
remaining part is goethite. Samples from the other field at the Gel- from the steel, the drills were cautiously examined under a stereo mi-
derner Fleuth show a more complex composition. Both of them consist croscope after cleaning for any damages to the drill or remnants of
mostly of carbonates (40–50%), and goethite (20–30%). Additionally sample material. Multiple samples were taken from all specimens to
glauconite from the bed rock is present in about 10% and 30%. Ad- investigate isotopic homogeneity. For samples ARC761 and ARC762
ditionally, the sample with the lower amount of glauconite also consists care was taken that each sampling spot was large enough to avoid
of about 10% of joosteite, an iron-manganese phosphate which might mixtures of optically different materials. Additionally to the spatially
have been introduced by the use of phosphate-rich fertilizers (e. g. resolved samples, several bigger pieces were detached, wrapped in
manure) or a prolonged use of this area as grazing land. paper and crushed for bulk analysis (Table 1). All other samples were

54
T. Rose et al. Journal of Archaeological Science 101 (2019) 52–62

Fig. 1. Location of the collected ore specimens from German sites analyzed in this study: (a) Eyller Bruch, and (b) Lahn-Dill ores from the vicinity of Wetzlar.
Additionally, the location area of (c) the Schwarzwald is indicated, where Markl et al. (2006) conducted their study (maps: openstreetmap.org).

spatially-resolved sampled to check for heterogeneity. 3.2.2. Mass spectrometry


Fe isotope analysis was carried out at the Laboratoire de Géologie de
Lyon, École Normale Supérieure de Lyon with the multi-collector
3.2. Fe isotope analysis plasma mass spectrometer (MC-ICP-MS) ThermoFisher Scientific
NeptunePlus in high resolution mode (Weyer and Schwieters, 2003).
3.2.1. Sample preparation Samples were dissolved on site in 2% HNO3 and further diluted with
Analytical work was carried out in a clean lab under a laminar flow 0.05 N HNO3. The MC-ICP-MS was equipped with Ni cones, a cyclonic
box at the Institut für Geowissenschaften, Goethe Universität Frankfurt. spray chamber and a Glass Expansion MicroMist nebulizer with an
Millipore® water with a resistivity of 18.2 MΩ cm, double distilled acids uptake rate of 100 μl/min. Uptake time for standard and samples was
and PTFE beakers were used throughout the analytical process. All 50 s and blanks were measured before each standard and sample with
samples were completely dissolved in a one-step acid digestion. Silicate 130 s uptake time and 180 s wash time. 1 ppm of IRMM-014 yielded an
containing samples were digested in ∼2 ml of 30 M HF and 6 M HNO3. intensity of ∼10 V on 56Fe. Mass bias and drift was corrected by
Silicate-free samples were digested with 1 ml of each, 6 M HCl and 6 M standard-sample-bracketing with the international reference material
HNO3. All samples were heated on a hot plate to 90 °C over-night (si- IRMM-014 (Craddock and Dauphas, 2011; Taylor et al., 1992) and
licates) or for several hours until complete dissolution was achieved doping with 1 ppm Ni (Maréchal et al., 1999; Poitrasson and Freydier,
(non-silicates). They were evaporated to dryness at the same tempera- 2005) prepared from the Alfa-Aesar Specpure® Ni plasma standard so-
ture, dissolved in 6 M HCl and dried down again. Finally, they were lution.
taken up in 0.5 ml 6 M HCl over-night for ion exchange chromato- Mass bias correction was carried out online during acquisition and δ
graphy (IEC). values were calculated in a spreadsheet according to the following
IEC was carried out according to an adapted protocol based on Sossi equation:
et al. (2015). Columns of 70 mm length with an inner diameter of 4 mm
and the resin AG1-X8 (100–200 mesh) were used. IEC induced Fe iso- ⎡ ( )F56 e
F54 e smp

tope fractionation was examined by splitting some samples before IEC δ56Fe [‰] = ⎢ − 1⎥*1000
⎢ ⎥
and proven to be absent. After IEC, samples were dried down at ∼70 °C.
They were re-dissolved in 1 ml 6 M HNO3 and evaporated to dryness


( )
F56 e
F54 e IRMM − 014 ⎥

again to remove remnants of resin.
Here, the index smp denotes the ratio 56Fe/54Fe of the sample and
the index IRMM-014 denotes the ratio of the international reference

55
T. Rose et al. Journal of Archaeological Science 101 (2019) 52–62

Fig. 2. Several nodules collected from the site Gelderner Fleuth, field1 (ore-field1) and Tote Rahm (ore-bog) as well as the specimen collected at Gelderner Fleuth,
field 2 (ore-field2) and polished sections of the ore from the Gelderner Fleuth used in the smelting experiments (ARC762) and the roasted ore (ARC761). The scales
for ore-field1, ore-field2, and ore-bog are the same. Also ARC761 and ARC762 have the same scale. Numbers in the big nodule of ore-bog as well as in ARC762
indicate sampling location of the respectively numbered sample.

material. The δ57Fe values are calculated accordingly. 4. Results


Each sample was run twice as duplicates and some samples were
repeatedly measured in different sessions to check for inter-run stabi- 4.1. Iron isotope analyses
lity. All data were further processed with R (R Core Team, 2017) in
RStudio® after acquisition. Outliers were identified and not regarded All data are given in Table 1. Data for the bog iron ore deposit
further, if their δ values derived from both isotope ratios differed more ranges between δ 56Fe = −1.29‰ and 1.16‰, which is nearly three
than the instrumental reproducibility of 0.045‰/amu between dupli- times larger than the range of the ores from the Montagne Noire given
cates and/or within the same analysis. The δ 56Fe values of all other in Milot (2016) and similar to the known overall range for hydro-
analyses were averaged per sample. Intensities of procedural blanks thermal ore deposits (2.40‰, cf. Fig. 6).
were more than 10000 times lower than sample intensities. The pre- Specimens collected at the Tote Rahm (ore-bog) gave isotopic
cision (external reproducibility) of the data is < 0.05‰ (2σ ), which is compositions in the range from δ 56Fe = −0.39‰ to 0.03‰ and scatter
comparable to previous studies (Albarède et al., 2011; Balter et al., around the clay sample ore-bog5, which yielded −0.23 ± 0.09‰.
2013). Compared to this sample, the clay found in a cavity (ore-bog4) is sig-
nificantly enriched in the heavier isotopes with a δ 56Fe value of 0.03 ±
0.09‰. A clear trend towards isotopically lighter values can be ob-
3.3. Loss on ignition served towards the crust and the soft ore of the same specimen (Fig. 3).
At the Gelderner Fleuth, the sandy ore (ore-field1) shows a similar
According to literature, the content of organic carbon (Ilina et al., variation and isotopic composition as the ones from Tote Rahm.
2013) and plants (Fantle and DePaolo, 2004; Wiederhold et al., 2006) However, the samples ore-field1-1 and ore-field1-2 are enriched in 54Fe
might have an influence on the iron isotopic composition of the bog and are more comparable with the specimen ore-field2. Interestingly,
iron ores. To take such a possible influence into account, loss on igni- specimen from both fields show contrasting trends between material
tion (LOI) was determined on samples from all bog iron ores to measure from the interior and the surface. While in ore-field1 the interior (ore-
the amount of organic matter. Carbonates were removed with 5% HCl field1-8) is slightly enriched in 54Fe and still within the range of the
previously to the temperature treatment. The samples were washed other specimens and the exterior of the same specimen (ore-field1-7),
several times in de-ionized water and dried at 105 °C for 18 h. Subse- the surface sampled as ore-field2-2 shows a marked depletion compared
quently the samples were ignited at 1050 °C to constant weight. to the other samples from this field. The latter observation is consistent

56
T. Rose et al. Journal of Archaeological Science 101 (2019) 52–62

Table 1
Fe isotope compositions of the analyzed samples. Analytical precision for all
samples is 0.09‰ for δ56Fe and 0.14‰ for δ57Fe. The last column (n) gives the
number of analyses per sample.
Sample Material δ56Fe [‰] δ57Fe [‰] n

Bog iron ores from the Eyller Bruch


Tote Rahm
ore-bog1 Ore nodule −0.11 −0.14 2
ore-bog2 Ore nodule −0.39 −0.57 5
ore-bog3 Ore nodule −0.22 −0.28 2
ore-bog4 Clay filling in cavity of nodule ore- 0.03 0.04 2
bog7
ore-bog5 Clay adhering to nodules −0.23 −0.37 3
ore-bog6 Crust around cavity in nodule ore- −0.11 −0.14 2
bog7
ore-bog7 Ore nodule −0.28 −0.38 3
Gelderner Fleuth, field 1 (sandy ore)
ore-field1-1 Ore nodule −0.70 −0.97 1
ore-field1-2 Ore nodule −0.75 −1.10 2
ore-field1-3 Ore nodule −0.36 −0.50 2
ore-field1-4 Ore nodule −0.09 −0.14 1
ore-field1-5 Ore nodule −0.30 −0.39 1
ore-field1-6 Ore nodule −0.31 −0.51 2
ore-field1-7 Ore nodule, same as ore-field1-8, −0.22 −0.32 2
from exterior areas
ore-field1-8 Ore nodule, same as ore-field1-7, −0.45 −0.71 3 Fig. 3. Fe isotope compositions of the bog iron ores from Eyller Bruch: Samples
from the interior from Tote Rahm (ore-bog) and Gelderner Fleuth, field 1 (ore-field1) represent
Gelderner Fleuth, field 2 different ore nodules except the ones additionally labeled on the right side of
ore-field2-2 Crust around surficial nodule −1.29 −1.94 2 the data points. These labeled samples belong to the same nodule per site and
ore-field2-3 Interior of surficial nodule −0.84 −1.23 3 the labels specify the sampling region of the respective nodule. For Gelderner
ore-field2-4 material from thicker crust −0.77 −1.16 6
Fleuth, field 2 (ore-field2) and the ore from the Gelderner Fleuth used for the
Gelderner Fleuth, ore used for smelting experiment
smelting experiment (ARC762) samples were taken from one specimen per
ARC762-1 Sampling location: see Fig. 2 0.03 0.02 2
ARC762-2 Encrusted root channel, see Fig. 2 1.16 1.70 2 deposit (cf. Table 1).
ARC762-3 Nodule-like piece 0.47 0.68 2
ARC762-4 Crust around big cavity 0.07 0.07 2
with isotope fractionation during weathering processes (e. g. Kiczka
ARC762-5 Bulk analysis 0.05 0.07 2
Materials from smelting experiment et al., 2011), but is an isolated case and thus will not be discussed
Roasted ore further.
ARC761-1 Loose material −0.08 −0.11 3 The ore used for the smelting experiment (ARC762) is slightly en-
ARC761-2 Piece of roasted ore −0.27 −0.42 4 riched in the heavier isotopes when the bulk sample ARC762-5 is
ARC761-3 Dark red material 0.27 0.37 2
ARC761-4 Light red material −0.14 −0.23 2
compared to the other ore specimens from the Gelderner Fleuth (Fig. 3).
ARC761-5 Piece of roasted ore 0.18 0.28 2 The sample ARC762-1 and the crust ARC762-4 also gave slightly po-
Bloom sitive δ 56Fe values of 0.03 ± 0.09‰ and 0.07 ± 0.09‰, respectively.
ARC759-1 −0.27 −0.42 2 In contrast to them the encrusted root channel ARC762-2 shows a
ARC759-2 −0.11 −0.23 2
pronounced shift of about one per mil towards higher δ 56Fe values and
ARC759-3 −0.12 −0.18 4
ARC759-4 −0.03 −0.05 2 sample ARC762-3 lies in-between with δ 56Fe = 0.47 ± 0.09‰
Slag (Fig. 3).
ARC760-1 −0.12 −0.23 2 In the roasted ore (ARC761) differences in the color of the sampled
ARC760-2 −0.14 −0.22 2 material also seem to indicate different isotopic compositions. The dark
ARC760-3 −0.16 −0.23 4
MAK-Sl1 −0.21 −0.35 1
red material ARC761-3 gave an δ 56Fe value of 0.27 ± 0.09‰ while
MAK-Sl2 −0.14 −0.20 2 the light red material ARC761-4 yielded δ 56Fe = −0.14 ± 0.09‰.
MAK-Sl3 −0.16 −0.22 2 The bulk samples ARC761-1 lies in the middle of them with an isotopic
Half product composition of δ 56Fe = −0.08 ± 0.09‰. Samples ARC761-5 and
MAK-Fe1 −0.05 −0.11 2
ARC761-2 are close to ARC761-3 and ARC761-4, respectively.
MAK-Fe2 −0.03 −0.04 2
MAK-Fe3 −0.05 −0.09 4 All other materials from the smelting experiment yielded widely
Lahn-Dill ores overlapping isotopic compositions (Fig. 4) and all of them overlap
Ore specimen from the Grube Fortuna within the analytical precision with the bulk samples of the ore and the
ore-GF1 0.61 0.89 2 roasted ore.
ore-GF2 0.51 0.74 2
ore-GF3 0.62 0.93 2
The Lahn-Dill-ores display a significantly heavier isotopic compo-
ore-GF4 0.55 0.86 2 sition than the bog iron ores (Table 1). All samples from ore-GF yielded
ore-GF5 0.53 0.78 2 isotopic compositions around δ 56Fe = 0.56‰. In contrast, samples
ore-GF6 0.52 0.79 2 from ore-W scatter between δ 56Fe = 0.49 ± 0.09‰ and 1.00 ± 0.09‰.
Ore specimen from ore heap near Wetzlar
The unaltered metallic core of the Roman semi-product SM9-
ore-W1 0.76 1.12 2
ore-W2 1.00 1.48 1 99.242. yielded an isotopic composition of δ 56Fe = −0.31 ± 0.09‰,
ore-W3 0.69 1.03 1 similar to δ 56Fe = −0.385‰, the value reported by Milot et al. (2016)
ore-W4 0.49 0.71 2 for the semi-product SM9-99.248 of the same wreck. It is slightly hea-
Les-Saintes-Maries-de-la-Mer, Roman semi product vier than the isotopic compositions of the other semi-products he re-
SM9-99.242.4 metallic core −0.31 −0.48 2
ports, but still within analytical precision.

57
T. Rose et al. Journal of Archaeological Science 101 (2019) 52–62

Fig. 4. Fe isotope compositions of the materials gained from the smelting ex-
periment with ore ARC762. In a polished section of the ore, a lot of plant-like
structures resembling cross-sections of stems and roots are visible. According to
literature, the considerably positive δ 56Fe values of the ore and the roasting ore
might be caused by a high content of organic matter, which is indicated by the
high loss on ignition of the ore (see chap. 5.2. for a detailed discussion).
Fig. 6. Fe isotope compositions for iron ores. Data from: Milot (2016) (Bassar,
Departement Tarn), Milot et al., (2016) (Montagne Noire), this study (Lahn-Dill
ore, Eyller Bruch), Cheng et al., (2015) (Gaosong), Markl et al., (2006)
(Schwarzwald). A compilation of the data can be found in the supplementary
data.

Table 2
Loss on ignition of the different ore samples.
Sample Material LOI (%)

ARC762-LOIc crusts 27.6 ± 2.3


ARC762-LOIs soft material 21.0 ± 0.8
ore-bog-LOI2 ore nodule 20.4 ± 1.0
ore-bog-LOI3 ore nodule 26.8 ± 0.5
ore-bog-LOI4 ore nodule 26.6 ± 1.0
ore-field1-LOI1 ore nodule 9.5 ± 0.8
ore-field1-LOI2 ore nodule 22.7 ± 1.8
ore-field1-LOI4 ore nodule 10.0 ± 0.7
ore-field1-LOI5 ore nodule 12.6 ± 0.4
ore-field1-LOI6 ore nodule 11.0 ± 1.0
ore-field1-LOI7 ore nodule 10.7 ± 3.9
ore-field2-LOIc crust around nodule 25.8 ± 0.9
ore-field2-LOIn complete nodule 28.9 ± 1.2

Fig. 5. The loss on ignition in per cent on carbonate-free samples from each
sampled bog iron ore site of the deposit Eyller Bruch. 2 and ARC762-3, which are strongly enriched in the heavier isotopes.
This observation coincides with current geochemical models of Fe
isotope fractionation during the precipitation of goethite in iron de-
4.2. Loss on ignition
posits. The heavier isotopes of dissolved reduced Fe in ionized form
preferentially oxidize and are adsorbed in the solid minerals, when iron
Results for the LOI on carbonate-free samples from bog iron ores of
(hydr)oxide minerals and goethite forms (Bullen et al., 2001; Frierdich
each site are shown in Fig. 5 and are given in Table 2. The ores ARC762,
et al., 2015; Mikutta et al., 2009). However, a very complex chemical
ore-bog and ore-field2 show loss on ignitions between 20 and 30%. In
mechanism provides permanent exchange of the isotopes between
contrast to them the quartz and feldspar rich samples from Gelderner
goethite and the dissolved iron (Beard et al., 2010; Handler et al., 2009;
Fleuth, field 1 (ore-field 1) only lost about 6 and 15% of their weight
Zarzycki and Rosso, 2017): Dissolved iron and goethite grains of small
except for one nodule (ore-field1-LOI2), which shows a similar weight
size (< 100 nm) had identical isotopic compositions after 30 days of
loss like the ores from the other sites.
equilibration (experiments of Handler et al., 2009), and extending this,
Handler et al. (2014) questioned a complete exchange for larger par-
5. Discussion ticles. Besides the grain sizes, the pH level of the solution might also
affect the exchange mechanism, although sufficiently long time scales
5.1. Fe isotope variation in bog iron ores are proposed to balance the deferring effect of a lower pH (Handler
et al., 2014; Reddy et al., 2015).
Isotopic variability of bog iron ores is very small within individual In the case of the Eyller Bruch, flow velocity of the ground water is
unaltered ore specimens. This is particularly true for the specimens low in both areas and the groundwater table is close enough to the
ARC762 and ore-field2, for which all unaltered samples of the two surface to ensure water saturation throughout most of the year. This
yielded nearly the same δ 56Fe value, and for the specimens of ore-bog. provides sufficient time for complete exchange between dissolved Fe
In ore-field1 most of the ore specimens show identical isotopic com- and iron (hydr)oxides at each site. Additionally, in the Schwarze Rahm
positions, only the samples ore-field1-1 and ore-field1-2 are sig- a thin film of iron (hydr)oxide particles was observed on the water
nificantly more enriched in 54Fe. Regarding the variation within the surface, indicating floating particles and, hence, undisturbed exchange
whole deposit, three groups are visible (Fig. 3). The isotopically light between the particles and the water.
group consists of the samples from ore-field2 as well as orefield1-1 and However, this mechanism does not account for the difference in the
ore-field1-2. Most other samples fall within the group with slightly Fe isotopic composition between the three groups. A possible ex-
negative values. The third group is established by the samples ARC762- planation for the more positive δ 56Fe values of ore-bog and ARC762 is

58
T. Rose et al. Journal of Archaeological Science 101 (2019) 52–62

the high amount of organic matter, indicated by their large LOI. There samples yielded identical isotopic compositions within analytical pre-
are multiple complementary explanations available: Fantle and cision. This strengthen the previous discussion in Milot (2016) and
DePaolo (2004) observed a preferential uptake of isotopically light iron Milot et al. (2016) concluding absence of Fe isotope fractionation from
by plants, leading to an enrichment of isotopically heavier dissolved ore to metal.
iron in the ground water. Ilina et al. (2013) concluded that a higher Nevertheless, deviations from the slightly negative δ 56Fe value of all
content of organic carbon preferentially enriches the heavier isotopes in smelting materials can be seen in the ore and the roasted ore and always
goethite due to a higher bond strength in bonds with carbon compared show isotopically heavier compositions. Because the whole ore is used,
to carbon-free bonds. Additionally, the lighter Fe isotopes are favored these heterogeneities will level each other out, resulting in the isotopic
during goethite dissolution with oxalate (Wiederhold et al., 2006). All composition of the bulk sample. Nonetheless, they have to be discussed.
the processes enriches the heavier isotopes in the mineral component. For the ore this was already done together with the other ore samples.
We are unable to differentiate from our data between these mechan- Concerning the roasted ore, the bulk analysis gave an isotopic compo-
isms, but it seems highly likely that one of these processes or a com- sition well within the variation of the other materials obtained in the
bination of them caused the slight but significant shift of ARC762 and smelting experiment. The sample from the dark red material is clearly
ore-bog towards heavier isotopic compositions compared with ore- isotopically heavier and the isotopic difference between the dark red
field2 and some samples from ore-field1. ARC762 in fact shows textural and the light red areas indicate isotopic heterogeneity within the
traces of plant material in the polished section and ferricreted root roasted ore. It seems very unlikely that the phase transformation from
channels. Concerning ore-bog, the survey area is densely covered with goethite to hematite during roasting (Rzepa et al., 2016) provoked an
plants and the soil is intensely rooted. isotopic fractionation through translocation. We therefore suggest that
The extent of such effects can be particularly well seen in ARC762-2 the heterogeneity is inherited from the bog iron ore. As was shown
and ARC762-3. The first one is an encrusted root channel and yielded above, a high content of organic matter seem to correlate with an iso-
the most positive isotopic composition of all samples analyzed. The topically heavier composition. The texture of the roasted ore is similar
latter does not show such a strongly positive δ 56Fe value, but is still well to the one of ARC762, indicating a formation in close contact to plants.
separated from all other ore samples, which might indicate an area with We therefore suggest that the different isotopic composition result from
an higher content of organic carbon than in the other samples. The different proportions of organic matter and that the dark red material
organic carbon content might also explain the intra-specimen differ- has a higher proportion of organic matter than the light red material,
ences between clay (ore-bog4) and surrounding iron (hydr)oxides (ore- resulting in a distinctly heavier isotopic composition.
bog7).
But two aspects seem to contradict this interpretation. Firstly, 5.3. Application potentials of Fe isotopes in archeometallurgy
samples from the specimen ore-field2 yielded a LOI as high as in the
ores ARC762 and ore-bog but show the isotopically lightest isotopic 5.3.1. Fe isotopes as provenance tracer for Fe artifacts
compositions of all ores. The combination of a high content of organic The Fe isotope compositions of bog iron ores from the Eyller Bruch
matter and light Fe isotope compositions indicate, that no noticeable ore field spread only over a per mil, if individual samples with ‘extreme’
influence of carbon containing bonds is present. The appearance of the isotopic compositions (ARC762-2, ARC762-3, ore-field2-2) are not
ore, nodules agglomerated by crusts, implies no or only a few plants taken into account (Fig. 6). For ARC762 it was shown in the previous
during the formation of this bog ore. Additionally, this is the only chapter that such strongly positive values like the ones from the both
sample containing a phosphate mineral. This mineral could point to- excluded data only occur at very small scales and have negligible in-
wards a prolonged use of the field as grazing land or the prolonged fluence on the bulk composition of the ore piece. And ore-field2-2
application of a phosphate rich fertilizer like manure. In both cases, big shows a strongly negative isotope composition due to weathering pro-
amounts of organic matter would have been introduced into the soil. cesses and therefore cannot be assumed as representative for the iso-
Therefore we suggest that organic matter in this ore specimen was topic composition of this bog iron ore.
mostly incorporated through remains of degraded or digested (plant) The spread of one per mil is similar to the variation of primary
remains. Such organic matter cannot remove the isotopically light iron hematite ore from the Schwarzwald region (Markl et al., 2006). It is
from the ground water and therefore would not result in isotopically considerably larger than the variation in ores from the Montagne Noire
heavy composition like it was the case for the other ores. and the Departement Tarn (Milot, 2016; Milot et al., 2016), and ores
The second aspect is the spread in the data from ore-field1. In this from the Gaosong deposit (Cheng et al., 2015).
ore, the influence of organic matter can only to some extent explain the The bog iron ores from the Eyller Bruch do not differ significantly in
spread in the data. The LOI of its samples has to be smaller than in all their isotope composition from the selected and compared deposits,
other samples because most of them are quartz and feldspar, which although the examples represent a wide variety of mineralization types
could not be removed. In this area of the deposit, the sand might also (primary and altered ores from different types of hydrothermal de-
have altered the isotopic composition of the Fe hydr (oxides). During posits, where the data from Markl et al. (2006) are from various de-
growth of goethite, silicates block some of the sites, on which dissolved posits in the Schwarzwald region). Based on their very limited variety
Fe ions adsorb. Therefore, growth of the mineral is limited but this also and significance, the reconstruction of ore provenance based on Fe
makes adsorption more selective for the heavier isotopes as they will isotopes alone would therefore be misleading, as the method has no
preferentially oxidize due to the stronger bonding environment discriminating potential as the geochronological isotope systems have,
(Wiederhold et al., 2007; Wu et al., 2011). Fractionation caused by this and as the available analyses have shown so far.
process does not seem as strong as the one caused by organic matter. An Even their use as a first approach as suggested by Milot et al. (2016)
intermediate isotopic composition, with respect to samples with low is not viable considering the large variety of ore types with comparable
and high content of organic matter might be expected, as it is the case and almost homogeneous isotopic compositions (Fig. 6). The applica-
in the data from the Eyller Bruch (Fig. 3). Further, the LOI of one ore tion could lead to matches without archeological sense. A good negative
nodule is as high as of the ores from other sites of the deposit. It must be example for such a misleading application would be the Eyller Bruch
assumed, that the isotopically heaviest compositions (esp. ore-field1-4) matching by analysis result with the Roman semi products of Les
results from a mixture of both processes. Saintes-Maries-de-la-Mer (SM9-99.242., this study; Milot et al., 2016;
Milot, 2016). This would clearly be a misjudgment as Baron et al.
5.2. Smelting experiment (2011) and Coustures et al. (2006) on the one hand conclude from other
methods than the iron isotopes that the iron for these objects originated
All materials from the smelting experiment or at least their bulk from the Montagne Noir and on the other hand no Roman mining

59
T. Rose et al. Journal of Archaeological Science 101 (2019) 52–62

activity is known for the Eyller Bruch. This is a strong example that Fe For all three aspects, the reconstruction of ore sources, the ex-
isotopes should not be used as a provenance tracer. This also holds true ploitation history of bog iron ore deposits, and the reconstruction of
on a more local scale as the strong overlap between the deposits from past landscapes it is highly desirable to have more iron isotope analyses
the Montage Noire and the Departement Tarn and from the different of iron ores from various deposits and mineralizations around the
mines from the Schwarzwald shows (Fig. 6, Supplement A1). world. In combination with a sound mineralogical description they are
One of the major roots on the Fe isotope composition of bog iron crucial to understand the influences on and the behavior of iron iso-
ores should be the iron source, whose isotopic signal was presumably topes within the different kinds of deposits and to draw further con-
altered during dissolution of the iron. Although no other isotopic data clusions about their potential for archeological research. Additionally,
than the results from this study are available for bog iron ores yet, it is other methods like trace elements should be used to reveal if there
thinkable that bog iron ores from other occurrences might have dif- might be other methods capable to differentiate between and/in within
ferent isotopic compositions, if their different iron sources does. As was bog iron ore deposits.
shown in the data from Eyller Bruch, also other influences might shift
the isotope signal towards heavier compositions (organic components,
pH-level, etc.). These mechanisms and the full isotopic range of bog 6. Conclusions
iron ores and other mineralizations must be investigated before Fe
isotope compositions of iron ores could again be discussed for use in Data from two sites of the historically mined bog iron ore deposit
archeometry. Eyller Bruch (Germany) provide an example for the influence of en-
As a positive result, one preliminary trend can be deduced. Ores vironmental parameters on the Fe isotope signature of iron ores and
from the Eyller Bruch, the Montagne Noire and the Departement Tarn suggest an intra-deposit zonation controlled by these factors. Materials
can be clearly separated from Lahn-Dill ores and ores from Bassar from a smelting experiment conducted with bog iron ore from this
(Fig. 6), with the latter two deposits showing a strongly positive iron deposit verify the conclusions of Milot (2016) and Milot et al. (2016)
isotope signature. Based on the present data it seems that very distinct that Fe isotopes do not fractionate in metallurgical processes. Even
ore formation processes and depositional influences might lead to dif- isotopic variation within the ore seems to become homogenized during
ferent iron isotope signatures in the ores. As the Lahn-Dill-study was a smelting since it was still detectable in the roasted ore but not in the
case study and does not represent the complete ore district, it cannot be bloom or slag.
assumed that the available data show the full isotopic spread of this ore The isotopic variation of the Eyller Bruch is comparable with the
type. Quartz veins in one of the ore pieces indicate hydrothermal al- range of other deposit types and largely overlaps with most of them;
teration and Markl et al. (2006) present evidence for Fe isotope frac- even with those that are entirely unrelated both geologically and re-
tionation through hydrothermal alteration. As the extent of hydro- gionally. Since Fe isotopes are sensitive to redox processes, this is to be
thermal alteration was not investigated, it must remain unclear whether expected. Although the scarcity of Fe isotope studies from iron ore
the measured isotopic composition represent unaltered or completely deposits other than banded iron formation impedes the assessment of Fe
altered ore or, most likely, something in between. isotopes as an additional tool rather than as being similar to the lead
isotope application in archeometry. The iron isotope system clearly
5.3.2. Intra-deposit variation as tool for environmental reconstructions? shows a lack in discriminatory power necessary for a widely applicable
Data from the Eyller Bruch indicate an influence of environmental provenance tracer, as it is predictable from geochemical knowledge.
parameters on the isotopic composition of bog iron ores (esp. content of Our results also emphasizes the importance of the historical and ar-
organic matter and silicates). This could point towards an isotopic cheological context for provenance reconstructions.
subdivision within bog iron ore deposits. Further studies should thus address the Fe isotopic variation of other
If a bog iron ore occurrence is identified by other methods as ori- bog iron ore deposits. If the variation turns out to be even larger than
ginal ore source, this might allow to narrow down the particular mining observed in this study this would finally fully exclude the use of Fe
area within a deposit, which was exploited during a particular time isotopes as a provenance tracer. Data from the Eyller Bruch provided
period. Such an approach would be similar to the use of copper isotopes positively first indications that Fe isotope compositions of bog ores can
to reconstruct the exploitation history of a deposit (Klein et al., 2010). be used for the reconstruction of past bog landscapes and the ex-
Another potential application of iron isotopes relates to intra-de- ploitation history of a mine and might even help to localize past mining
posit variation. As they are influenced by environmental parameters, areas.
bog iron ores might yield some information about the appearance of the
bog iron ore deposit or its underlying substrate even if its location can
no longer be reconstructed. Consequently, they might contribute to the Acknowledgements
reconstruction of past landscapes.
Admittedly, such information might be gathered only from the ore The Deutsches Bergbau-Museum Bochum and the Institut für
itself and not from metallurgical by-products or the metal. Besides the Geowissenschaften of the Goethe Universität Frankfurt are thanked for
general positive inference, that the overall Fe isotope composition is not funding the analytical work of Thomas Rose's Master thesis. The
altered during smelting, isotopic differences within the ore seem to be Hermann-Willkomm-Stiftung thankfully financially supported the
homogenized and information about the deposit might be erased in the travel of T. Rose to Lyon. We thank Luc Long for the allowance to
objects. There is one chance for use if pronounced intra-deposit zona- sample and analyze the Roman semi-product. T. Rose is grateful for the
tion occurs and ores from different zones are not mixed but result in support of Maria Arians-Kronenberg and Guntram Gassmann during
isotopically distinct furnace charges. Subsequently isotopically distinct fieldwork. We thank our colleagues Katrin Westner and Eveline
metal objects from the different zones could result, which were in- Salzmann for fruitful discussions and Chiara Girotto for improvement of
formative for the deposit. If they are previously provenanced by other the language. The helpful comments of the anonymous reviewers are
methods such as trace elements or osmium isotopes, the structure of the gratefully acknowledged.
bog can be reconstructed. Or vice versa, combining Fe isotope data
from local bog ores found in the archeological record with methods for
landscape reconstruction might allow to infer the location of past Appendix A. Supplementary data
mining areas. To implement these applications, future studies must
confirm such distinctive intra-deposit variations and must link specific Supplementary data related to this article can be found at https://
isotope fractionations to distinct environmental parameters. doi.org/10.1016/j.jas.2018.11.005.

60
T. Rose et al. Journal of Archaeological Science 101 (2019) 52–62

References 12.022.
Dillmann, P., Schwab, R., Bauvais, S., Brauns, M., Disser, A., Leroy, S., Gassmann, G.,
Fluzin, P., 2017. Circulation of iron products in the North-alpine area during the end
Albarède, F., 2015. Metal stable isotopes in the human body: a tribute of geochemistry to of the first Iron Age (6th -5th c. BC): a combination of chemical and isotopic ap-
medicine. Elements 11, 265–269. https://doi.org/10.2113/gselements.11.4.265. proaches. J. Archaeol. Sci. 87, 108–124. https://doi.org/10.1016/j.jas.2017.10.002.
Albarède, F., Télouk, P., Lamboux, A., Jaouen, K., Balter, V., 2011. Isotopic evidence of Disser, A., Dillmann, P., Bourgain, C., L’Héritier, M., Vega, E., Bauvais, S., Leroy, M.,
unaccounted for Fe and Cu erythropoietic pathways. Metall 3, 926–933. https://doi. 2014. Iron reinforcements in Beauvais and Metz cathedrals: from bloomery or finery?
org/10.1039/c1mt00025j. The use of logistic regression for differentiating smelting processes. J. Archaeol. Sci.
Balter, V., Lamboux, A., Zazzo, A., Télouk, P., Leverrier, Y., Marvel, J., Moloney, A.P., 42, 315–333. https://doi.org/10.1016/j.jas.2013.10.034.
Monahan, F.J., Schmidt, O., Albarède, F., 2013. Contrasting Cu, Fe, and Zn isotopic Eerkens, J.W., Barfod, G.H., Vaughn, K.J., Williams, P.R., Lesher, C.E., 2014. Iron isotope
patterns in organs and body fluids of mice and sheep, with emphasis on cellular analysis of red and black pigments on pottery in Nasca, Peru. Archaeological and
fractionation. Metall 5, 1470–1482. https://doi.org/10.1039/c3mt00151b. Anthropological Sciences 6, 241–254. https://doi.org/10.1007/s12520-013-0151-6.
Banning, A., 2008. Bog iron ores and their potential role in arsenic dynamics: an overview Fantle, M.S., DePaolo, D.J., 2004. Iron isotopic fractionation during continental weath-
and a “paleo example”. Eng. Life Sci. 8, 641–649. https://doi.org/10.1002/elsc. ering. Earth Planet Sci. Lett. 228, 547–562. https://doi.org/10.1016/j.epsl.2004.10.
200800014. 013.
Baron, S., Coustures, M.-P., Béziat, D., Guérin, M., Huez, J., Robbiola, L., 2011. Lingots de Frierdich, A.J., Beard, B.L., Rosso, K.M., Scherer, M.M., Spicuzza, M.J., Valley, J.W.,
plomb et barres de fer des épaves romaines des Saintes-maries-de-la-mer (Bouches- Johnson, C.M., 2015. Low temperature, non-stoichiometric oxygen-isotope exchange
du-Rhône, France): questions de traçabilité comparée: questions de traçabilité coupled to Fe(II)–goethite interactions. Geochem. Cosmochim. Acta 160, 38–54.
comparée. Rev. Archéol. Narbonnaise 44, 71–98. https://doi.org/10.1016/j.gca.2015.03.029.
Beard, B.L., Handler, R.M., Scherer, M.M., Wu, L., Czaja, A.D., Heimann, A., Johnson, Frierdich, A.J., Beard, B.L., Scherer, M.M., Johnson, C.M., 2014. Determination of the Fe
C.M., 2010. Iron isotope fractionation between aqueous ferrous iron and goethite. (II)aq–magnetite equilibrium iron isotope fractionation factor using the three-isotope
Earth Planet Sci. Lett. 295, 241–250. https://doi.org/10.1016/j.epsl.2010.04.006. method and a multi-direction approach to equilibrium. Earth Planet Sci. Lett. 391,
Beard, B.L., Johnson, C.M., 1999. High precision iron isotope measurements of terrestrial 77–86. https://doi.org/10.1016/j.epsl.2014.01.032.
and lunar materials. Geochem. Cosmochim. Acta 63, 1653–1660. https://doi.org/10. Halkon, P., 2014. Iron and the Parisi - socio-economic and ritual aspects of the iron in-
1016/S0016-7037(99)00089-7. dustry in roman east Yorkshire, UK. In: Cech, B., Rehren, T. (Eds.), Early Iron in
Beard, B.L., Johnson, C.M., Damm, K.L., von Poulson, R.L., 2003. Iron isotope constraints Europe, Monographies Instrumentum. Editions Monique Mergoil, Montagnac, pp.
on Fe cycling and mass balance in oxygenated earth oceans. Geology 31 (629) 203–214.
https://doi.org/10.1130/0091-7613(2003)031 < 0629:IICOFC > 2.0.CO;2. Halverson, G.P., Poitrasson, F., Hoffman, P.F., Nédélec, A., Montel, J.-M., Kirby, J., 2011.
Benvenuti, M., Orlando, A., Borrini, D., Chiarantini, L., Costagliola, P., Mazzotta, C., Fe isotope and trace element geochemistry of the neoproterozoic syn-glacial Rapitan
Rimondi, V., 2016. Experimental smelting of iron ores from Elba island (Tuscany, iron formation. Earth Planet Sci. Lett. 309, 100–112. https://doi.org/10.1016/j.epsl.
Italy): results and implications for the reconstruction of ancient metallurgical pro- 2011.06.021.
cesses and iron provenance. J. Archaeol. Sci. 70, 1–14. https://doi.org/10.1016/j.jas. Handler, R.M., Beard, B.L., Johnson, C.M., Scherer, M.M., 2009. Atom exchange between
2016.04.008. aqueous Fe(II) and goethite: an Fe isotope tracer study. Environ. Sci. Technol. 43,
Blakelock, E., Martinón-Torres, M., Veldhuijzen, H.A., Young, T., 2009. Slag inclusions in 1102–1107. https://doi.org/10.1021/es802402m.
iron objects and the quest for provenance: an experiment and a case study. J. Handler, R.M., Frierdich, A.J., Johnson, C.M., Rosso, K.M., Beard, B.L., Wang, C., Latta,
Archaeol. Sci. 36 (8), 1745–1757. https://doi.org/10.1016/j.jas.2009.03.032. D.E., Neumann, A., Pasakarnis, T., Premaratne, W.A.P.J., Scherer, M.M., 2014. Fe(II)-
Brauns, M., Schwab, R., Gassmann, G., Wieland, G., Pernicka, E., 2013. Provenance of catalyzed recrystallization of goethite revisited. Environ. Sci. Technol. 48,
iron age iron in Southern Germany: a new approach. J. Archaeol. Sci. 40, 841–849. 11302–11311. https://doi.org/10.1021/es503084u.
https://doi.org/10.1016/j.jas.2012.08.044. Hansel, C.M., Benner, S.G., Fendorf, S., 2005. Competing Fe(II)-induced mineralization
Bullen, T.D., White, A.F., Childs, C.W., Vivit, D.V., Schulz, M.S., 2001. Demonstration of pathways of ferrihydrite. Environ. Sci. Technol. 39, 7147–7153. https://doi.org/10.
significant abiotic iron isotope fractionation in nature. Geology 29 (699) https:// 1021/es050666z.
doi.org/10.1130/0091-7613(2001)029 < 0699:DOSAII > 2.0.CO;2. Hoefs, J., 2009. Stable Isotope Geochemistry. Springer, Berlin; Heidelberg. https://doi.
Charlton, M.F., 2015. The last frontier in “sourcing”: the hopes, constraints and future for org/10.1007/978-3-540-70708-0.
iron provenance research. J. Archaeol. Sci. 56, 210–220. https://doi.org/10.1016/j. Horn, I., Blanckenburg, F., von Schoenberg, R., Steinhoefel, G., Markl, G., 2006. In situ
jas.2015.02.017. iron isotope ratio determination using UV-femtosecond laser ablation with applica-
Cheng, Y., Mao, J., Zhu, X., Wang, Y., 2015. Iron isotope fractionation during supergene tion to hydrothermal ore formation processes. Geochem. Cosmochim. Acta 70,
weathering process and its application to constrain ore genesis in Gaosong deposit, 3677–3688. https://doi.org/10.1016/j.gca.2006.05.002.
Gejiu district, SW China. Gondwana Res. 27, 1283–1291. https://doi.org/10.1016/j. Ilina, S.M., Poitrasson, F., Lapitskiy, S.A., Alekhin, Y.V., Viers, J., Pokrovsky, O.S., 2013.
gr.2013.12.006. Extreme iron isotope fractionation between colloids and particles of boreal and
Coustures, M.P., Béziat, D., Tollon, F., Domergue, C., Long, L., Rebiscoul, A., 2003. The temperate organic-rich waters. Geochem. Cosmochim. Acta 101, 96–111. https://doi.
use of trace element analysis of entrapped slag inclusions to establish ore – bar iron org/10.1016/j.gca.2012.10.023.
links: examples from two gallo-roman iron-making sites in France (Les Martys, Jang, J.-H., Mathur, R., Liermann, L.J., Ruebush, S., Brantley, S.L., 2008. An iron isotope
Montagne Noire, and Les Ferrys, Loiret). Archaeometry 45 (4), 599–613. https://doi. signature related to electron transfer between aqueous ferrous iron and goethite.
org/10.1046/j.1475-4754.2003.00131.x. Chem. Geol. 250, 40–48. https://doi.org/10.1016/j.chemgeo.2008.02.002.
Coustures, M.-P., Rico, C., Béziat, D., Djaoui, D., Long, L., Domergue, C., Tollon, F., 2006. Joosten, I., Jansen, J., Kars, H., 1998. Geochemistry and the past: estimation of the output
La provenance des barres de fer romaines des Saintes-maries-de-la-mer (Bouches-du- of a Germanic iron production site in The Netherlands. J. Geochem. Explor. 62,
Rhône): Étude archéologique et archéométrique. Gallia 63, 243–261. https://doi.org/ 129–137. https://doi.org/10.1016/S0375-6742(97)00043-5.
10.3406/galia.2006.3297. Kaczorek, D., Sommer, M., 2003. Micromorphology, chemistry, and mineralogy of bog
Craddock, P.R., Dauphas, N., 2011. Iron isotopic compositions of geological reference iron ores from Poland. Catena 54, 393–402. https://doi.org/10.1016/S0341-
materials and chondrites. Geostand. Geoanal. Res. 35, 101–123. https://doi.org/10. 8162(03)00133-4.
1111/j.1751-908X.2010.00085.x. Kiczka, M., Wiederhold, J.G., Frommer, J., Voegelin, A., Kraemer, S.M., Bourdon, B.,
Crerar, D., Knox, G., Means, J., 1979. Biogeochemistry of bog iron in the New Jersey pine Kretzschmar, R., 2011. Iron speciation and isotope fractionation during silicate
barrens. Chem. Geol. 24, 111–135. https://doi.org/10.1016/0009-2541(79)90016-0. weathering and soil formation in an alpine glacier forefield chronosequence.
Dassel, W., 2007. Beiträge zur Geologie, Archäologie und Geschichte an Rhein und Maas, Geochem. Cosmochim. Acta 75, 5559–5573. https://doi.org/10.1016/j.gca.2011.07.
Veröffentlichungen des historischen Vereins für Geldern und Umgegend. Historischer 008.
Verein für Geldern und Umgegend, Geldern. Klein, S., Brey, G.P., Durali-Müller, S., Lahaye, Y., 2010. Characterisation of the raw metal
Dauphas, N., John, S.G., Rouxel, O.J., 2017. Iron isotope systematics. Rev. Mineral. sources used for the production of copper and copper-based objects with copper
Geochem. 82, 415–510. https://doi.org/10.2138/rmg.2017.82.11. isotopes. Archaeological and Anthropological Sciences 2, 45–56. https://doi.org/10.
Dauphas, N., Rouxel, O.J., 2006. Mass spectrometry and natural variations of iron iso- 1007/s12520-010-0027-y.
topes. Mass Spectrom. Rev. 25, 515–550. https://doi.org/10.1002/mas.20078. Larsen, J.H., Rundberget, B., 2014. Iron bloomery in south and central Norway, 300 BC -
Degryse, P., Schneider, J., Kellens, N., Waelkens, M., Muchez, P., 2007. Tracing the re- 500 AD. In: Cech, B., Rehren, T. (Eds.), Early Iron in Europe, Monographies
sources of iron working at ancient Sagalassos (South-West Turkey): a combined lead Instrumentum. Editions Monique Mergoil, Montagnac, pp. 231–248.
and strontium isotope study on iron artifacts and ores. Archaeometry 49, 75–86. Leroy, S., Cohen, S.X., Verna, C., Gratuze, B., Téreygeol, F., Fluzin, P., Bertrand, L.,
https://doi.org/10.1111/j.1475-4754.2007.00288.x. Dillmann, P., 2012. The medieval iron market in Ariège (France). Multidisciplinary
Degryse, P., Schneider, J.C., Muchez, P., 2009. Combined Pb–Sr isotopic analysis in analytical approach and multivariate analyses. J. Archaeol. Sci. 39 (4), 1080–1093.
provenancing late roman iron raw materials in the territory of Sagalassos (SW https://doi.org/10.1016/j.jas.2011.11.025.
Turkey). Archaeological and Anthropological Sciences 1, 155–159. https://doi.org/ L’Héritier, M., Leroy, S., Dillmann, P., Gratuze, B., Dussubieux, L., Golitko, M., Gratuze,
10.1007/s12520-009-0010-7. B., 2016. Characterization of slag inclusions in iron objects. In: Recent Advances in
Desaulty, A.-M., Dillmann, P., L’Héritier, M., Mariet, C., Gratuze, B., Joron, J.-L., Fluzin, Laser Ablation ICP-MS for Archaeology. Springer, Heidelberg, Berlin, pp. 213–228.
P., 2009. Does it come from the Pays de Bray?: examination of an origin hypothesis https://doi.org/10.1007/978-3-662-49894-1_14.
for the ferrous reinforcements used in French medieval churches using major and Long, L., Rico, C., Domergue, C., 2002. Les épaves antiques de Camargue et le commerce
trace element analyses. J. Archaeol. Sci. 36 (10), 2445–2462. https://doi.org/10. maritime du fer en Méditerranée nord-occidentale (Ier siècle avant J.-C. - Ier siècle
1016/j.jas.2009.07.002. après J.-C.). In: Khanoussi, M. (Ed.), L’ Africa Romana, Pubblicazioni Del
Dillmann, P., L’Héritier, M., 2007. Slag inclusion analyses for studying ferrous alloys Dipartimento Di Storia Dell'Universitá Degli Studi Di Sassari. Carocci, Roma, pp.
employed in French medieval buildings: supply of materials and diffusion of smelting 161–188.
processes. J. Archaeol. Sci. 34 (11), 1810–1823. https://doi.org/10.1016/j.jas.2006. Maréchal, C.N., Télouk, P., Albarède, F., 1999. Precise analysis of copper and zinc isotopic

61
T. Rose et al. Journal of Archaeological Science 101 (2019) 52–62

compositions by plasma-source mass spectrometry. Chem. Geol. 156, 251–273. Schwertmann, U., Friedl, J., Stanjek, H., 1999. From Fe(III) ions to ferrihydrite and then
https://doi.org/10.1016/S0009-2541(98)00191-0. to hematite. J. Colloid Interface Sci. 209, 215–223. https://doi.org/10.1006/jcis.
Markl, G., Blanckenburg, F. von, Wagner, T., 2006. Iron isotope fractionation during 1998.5899.
hydrothermal ore deposition and alteration. Geochem. Cosmochim. Acta 70, Schwertmann, U., Murad, E., 1983. Effect of pH on the formation of goethite and hematite
3011–3030. https://doi.org/10.1016/j.gca.2006.02.028. from ferrihydrite. Clay Clay Miner. 31, 277–284.
Mathur, R., Fantle, M.S., 2015. Copper isotopic perspectives on supergene processes: Sommer, H., 1963. Erzsuche im Eyller Bruch. In: Geldrischer Heimatkalender 1963, pp.
implications for the global Cu cycle. Elements 11 (5), 323–329. https://doi.org/10. 112–113.
2113/gselements.11.5.323. Sossi, P.A., Halverson, G.P., Nebel, O., Eggins, S.M., 2015. Combined separation of Cu, Fe
Mikutta, C., Wiederhold, J.G., Cirpka, O.A., Hofstetter, T.B., Bourdon, B., Gunten, U.V., and Zn from rock matrices and improved analytical protocols for stable isotope de-
2009. Iron isotope fractionation and atom exchange during sorption of ferrous iron to termination. Geostand. Geoanal. Res. 39, 129–149. https://doi.org/10.1111/j.1751-
mineral surfaces. Geochem. Cosmochim. Acta 73, 1795–1812. https://doi.org/10. 908X.2014.00298.x.
1016/j.gca.2009.01.014. Steinhoefel, G., Horn, I., Blanckenburg, F. von, 2009. Micro-scale tracing of Fe and Si
Milot, J., 2016. Utilisation des isotopes du fer pour le traçage des métaux ancien: isotope signatures in banded iron formation using femtosecond laser ablation.
Développement méthodologique et applications archéologiques (PhD thesis). Geochem. Cosmochim. Acta 73, 5343–5360. https://doi.org/10.1016/j.gca.2009.05.
Université Toulouse 3 Paul Sabatier. Sciences de la Terre et des Planètes Solides, 037.
Toulouse. Taylor, P.D.P., Maeck, R., Bièvre, P. de, 1992. Determination of the absolute isotopic
Milot, J., Poitrasson, F., Baron, S., Coustures, M.-P., 2016. Iron isotopes as a potential tool composition and atomic weight of a reference sample of natural iron. Int. J. Mass
for ancient iron metals tracing. J. Archaeol. Sci. 76, 9–20. https://doi.org/10.1016/j. Spectrom. Ion Process. 121, 111–125. https://doi.org/10.1016/0168-1176(92)
jas.2016.10.003. 80075-C.
Moynier, F., Vance, D., Fujii, T., Savage, P., 2017. The isotope geochemistry of zinc and Teutsch, N., Gunten, U.V., Porcelli, D., Cirpka, O.A., Halliday, A.N., 2005. Adsorption as a
copper. Rev. Mineral. Geochem. 82 (1), 543–600. https://doi.org/10.2138/rmg. cause for iron isotope fractionation in reduced groundwater. Geochem. Cosmochim.
2017.82.13. Acta 69, 4175–4185. https://doi.org/10.1016/j.gca.2005.04.007.
Pagès, G., 2008. La métallurgie du fer en france méditerranéenne de l'Antiquité au début Thelemann, M., Bebermeier, W., Hoelzmann, P., Lehnhardt, E., 2017. Bog iron ore as a
du moyen Âge: Jalons d’une approche interdisciplinaire (PhD thesis). Université resource for prehistoric iron production in central Europe — a case study of the
Monpellier III - Paul Valéry; Arts er Lettres. Langues et Sciences Humaines et Sociales, Widawa catchment area in eastern Silesia, Poland. Catena 149, 474–490. https://doi.
Montpellier. org/10.1016/j.catena.2016.04.002.
Pagès, G., Long, L., Fluzin, P., Dillmann, P., 2008. Réseaux de production et standards de Toby, B.H., Dreele, R.B. von, 2013. GSAS-II: the genesis of a modern open-source all
commercialisation du fer antique en Méditerranée: Les demi-produits des épaves purpose crystallography software package. J. Appl. Crystallogr. 46 (2), 544–549.
romaines des Saintes-maries-de-la-mer (Bouches-du-Rhône). Rev. Archéol. https://doi.org/10.1107/S0021889813003531.
Narbonnaise 41, 261–283. https://doi.org/10.3406/ran.2008.1194. Wang, Y., Zhu, X.-k., Cheng, Y., 2015. Fe isotope behaviours during sulfide-dominated
Pedersen, H.D., Postma, D., Jakobsen, R., Larsen, O., 2005. Fast transformation of iron skarn-type mineralisation. J. Asian Earth Sci. 103, 374–392. https://doi.org/10.
oxyhydroxides by the catalytic action of aqueous Fe(II). Geochem. Cosmochim. Acta 1016/j.jseaes.2014.11.005.
69, 3967–3977. https://doi.org/10.1016/j.gca.2005.03.016. Wawryk, C.M., Foden, J.D., 2015. Fe-isotope fractionation in magmatic-hydrothermal
Poage, M.A., Sjostrom, D.J., Goldberg, J., Chamberlain, C., Furniss, G., 2000. Isotopic mineral deposits: a case study from the Renison Sn–W deposit, Tasmania. Geochem.
evidence for Holocene climate change in the Northern Rockies from a goethite-rich Cosmochim. Acta 150, 285–298. https://doi.org/10.1016/j.gca.2014.09.044.
ferricrete chronosequence. Chem. Geol. 166, 327–340. https://doi.org/10.1016/ Weyer, S., Schwieters, J.B., 2003. High precision Fe isotope measurements with high mass
S0009-2541(99)00220-X. resolution MC-ICPMS. Int. J. Mass Spectrom. 226, 355–368. https://doi.org/10.
Poitrasson, F., 2006. On the iron isotope homogeneity level of the continental crust. 1016/S1387-3806(03)00078-2.
Chem. Geol. 235, 195–200. https://doi.org/10.1016/j.chemgeo.2006.06.010. Wiederhold, J.G., Kraemer, S.M., Teutsch, N., Borer, P.M., Halliday, A.N., Kretzschmar,
Poitrasson, F., Freydier, R., 2005. Heavy iron isotope composition of granites determined R., 2006. Iron isotope fractionation during proton-promoted, ligand-controlled, and
by high resolution MC-ICP-MS. Chem. Geol. 222 (1–2), 132–147. https://doi.org/10. reductive dissolution of goethite. Environ. Sci. Technol. 40, 3787–3793. https://doi.
1016/j.chemgeo.2005.07.005. org/10.1021/es052228y.
Poulson, R.L., Johnson, C.M., Beard, B.L., 2005. Iron isotope exchange kinetics at the Wiederhold, J.G., Teutsch, N., Kraemer, S.M., Halliday, A.N., Kretzschmar, R., 2007. Iron
nanoparticulate ferrihydrite surface. Am. Mineral. 90, 758–763. https://doi.org/10. isotope fractionation in oxic soils by mineral weathering and podzolization.
2138/am.2005.1802. Geochem. Cosmochim. Acta 71, 5821–5833. https://doi.org/10.1016/j.gca.2007.07.
R Core Team, 2017. R: a Language and Environment for Statistical Computing. 023.
Reddy, T.R., Frierdich, A.J., Beard, B.L., Johnson, C.M., 2015. The effect of pH on stable Wu, L., Beard, B.L., Roden, E.E., Johnson, C.M., 2011. Stable iron isotope fractionation
iron isotope exchange and fractionation between aqueous Fe(II) and goethite. Chem. between aqueous Fe(II) and hydrous ferric oxide. Environ. Sci. Technol. 45,
Geol. 397, 118–127. https://doi.org/10.1016/j.chemgeo.2015.01.018. 1847–1852. https://doi.org/10.1021/es103171x.
Rzepa, G., Bajda, T., Gaweł, A., Debiec, K., Drewniak, L., 2016. Mineral transformations Yapp, C.J., 2001. Rusty relics of earth history: iron(III) oxides, isotopes, and surficial
and textural evolution during roasting of bog iron ores. J. Therm. Anal. Calorim. 123, environments. Annu. Rev. Earth Planet Sci. 29, 165–199. https://doi.org/10.1146/
615–630. https://doi.org/10.1007/s10973-015-4925-1. annurev.earth.29.1.165.
Schwab, R., Heger, D., Höppner, B., Pernicka, E., 2006. The provenance of iron artifacts Zarzycki, P., Rosso, K.M., 2017. Stochastic simulation of isotopic exchange mechanisms
from Manching: a multi-technique approach. Archaeometry 48, 433–452. https:// for Fe(II)-catalyzed recrystallization of goethite. Environ. Sci. Technol. 51,
doi.org/10.1111/j.1475-4754.2006.00265.x. 7552–7559. https://doi.org/10.1021/acs.est.7b01491.

62

You might also like