You are on page 1of 20

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/329783054

Azania: Archaeological Research in Africa Copper wire objects from Jahunda


and Little Mapela: technology, value systems and networks in Iron Age
southern Africa

Preprint · November 2018


DOI: 10.1080/0067270X.2018.1540213

CITATION READS

1 311

3 authors:

Foreman Bandama Munyaradzi Manyanga


Sol Plaatje University, Kimberley, South africa Great Zimbabwe University
23 PUBLICATIONS   229 CITATIONS    41 PUBLICATIONS   472 CITATIONS   

SEE PROFILE SEE PROFILE

Shadreck Chirikure
University of Cape Town
102 PUBLICATIONS   1,268 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Land use and land cover in southern Africa from 6000 years ago to the present View project

Resilience, livelihoods and traditions in the Shashi-Limpopo Basin View project

All content following this page was uploaded by Foreman Bandama on 22 December 2018.

The user has requested enhancement of the downloaded file.


Azania: Archaeological Research in Africa

ISSN: 0067-270X (Print) 1945-5534 (Online) Journal homepage: http://www.tandfonline.com/loi/raza20

Copper wire objects from Jahunda and Little


Mapela: technology, value systems and networks
in Iron Age southern Africa

Foreman Bandama, Munyaradzi Manyanga & Shadreck Chirikure

To cite this article: Foreman Bandama, Munyaradzi Manyanga & Shadreck Chirikure
(2018): Copper wire objects from Jahunda and Little Mapela: technology, value systems
and networks in Iron Age southern Africa, Azania: Archaeological Research in Africa, DOI:
10.1080/0067270X.2018.1540213

To link to this article: https://doi.org/10.1080/0067270X.2018.1540213

Published online: 27 Nov 2018.

Submit your article to this journal

View Crossmark data

Citing articles: 1 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=raza20
AZANIA: ARCHAEOLOGICAL RESEARCH IN AFRICA
https://doi.org/10.1080/0067270X.2018.1540213

Copper wire objects from Jahunda and Little Mapela:


technology, value systems and networks in Iron Age southern
Africa
a,b
Foreman Bandama , Munyaradzi Manyangab,c and Shadreck Chirikureb
a
Heritage Department, School of Humanities, Sol Plaatje University, Kimberley, 8301, South Africa;; bMaterials
Laboratory, Department of Archaeology, University of Cape Town, Rondebosch, 7700, South Africa;;
c
Department of History, University of Zimbabwe, Mount Pleasant, Harare, Zimbabwe

ABSTRACT ARTICLE HISTORY


The present study explores the archaeometallurgy of flexible Received 14 August 2018
cuprous wound wires recovered from two precolonial Shona Accepted 1 October 2018
centres of power, Jahunda and Little Mapela, separated by 60
KEYWORDS
kilometres in Gwanda, southwestern Zimbabwe. Widely believed Copper wire objects; value
to have been fashion accessories, archaeologists have traditionally systems; networks; state
consigned such objects to appendices in site reports and other formation; southern Africa;
publications. Chronological and technological information Iron Age
gathered from this category of objects indicates that Jahunda
exploited more tin bronzes than pure, unalloyed copper. The
converse holds for Little Mapela. However, the presence of copper
and its alloys brass (with zinc) and bronze (with tin), in a region
with no known sources of any of the raw materials, poses
important questions about the role and value of these objects in
both their immediate societies and networked regional places.
This study shows that as metal objects, flexible wound wires were
an important component of the political economy and conjoined
diverse worlds — healing, spiritual, decorative, technological and
much more — that made their production, circulation and use a
microcosm of resilient networks and changing societal values.

RESUMÉ
Cette étude explore l’archéométallurgie des fils cuivreux flexibles
enroulés retrouvés sur deux centres de pouvoir Shona
précoloniaux, à savoir Jahunda et Little Mapela. Ces sites, dans le
Gwanda, sud-ouest du Zimbabwe, sont distants de 60 kilomètres
l’un de l’autre. Les objets en question ont typiquement été
interprétés comme des accessoires de mode, et les archéologues
les ont traditionnellement relégués à des annexes dans les
rapports de site et autres publications. Les informations
chronologiques et technologiques recueillies à la base de cette
catégorie d’objets indiquent que Jahunda exploita davantage le
bronze d’étain que le cuivre pur non allié. L’inverse est vrai pour
Little Mapela. Cependant, la présence de cuivre et de ses alliages
— le laiton (avec le zinc) et le bronze (avec l’étain) — dans une
région apparemment dépourvue de toute source de ces matières
premières pose des questions importantes quant au rôle et à la
valeur de ces objets dans ces sociétés et dans les lieux qui y

CONTACT Foreman Bandama fbandama@gmail.com


© 2018 Informa UK Limited, trading as Taylor & Francis Group
2 F. BANDAMA ET AL.

étaient connectés. Cette étude montre que les fils souples enroulés
constituaient, en tant qu’objets métalliques, un élément important
de l’économie politique et réunissaient des univers divers —
guérison, spirituel, décoration, technologique et bien d’autres —
qui faisaient de leur production, circulation et usage un
microcosme de réseaux résilients et de valeurs sociétales
changeantes.

Introduction
Copper metallurgy was introduced to southern Africa (Figure 1) alongside iron by Bantu
agriculturalists early in the first millennium AD. The late first and early second millennia
also saw the introduction of tin (alloyed with copper to produce bronze) and ready made
brass (an alloy of copper and zinc) (Bandama et al. 2017). Initially worked on a very small
scale, the exploitation of copper increased in intensity and concentration as populations
and interaction networks expanded over the course of the first millennium (Thondhlana
2013). Some of the sites with the earliest evidence of copper metallurgy include Kansanshi

Figure 1. Southern Africa showing Little Mapela, Jahunda and other regional sites with cuprous flexible
wound wires.
AZANIA: ARCHAEOLOGICAL RESEARCH IN AFRICA 3

in Zambia (c. cal. AD 450) and Broederstroom in South Africa (c. cal. AD 500) (Miller and
van der Merwe 1994; Bisson 2000; Huffman 2007) (Figure 1). Available evidence shows
that copper was rarely used to make heavy duty utilitarian objects. Rather, it was manipu-
lated to produce beads and other decorative items. By the late first millennium AD, copper
decorative items including drawn and hammered wires became plentiful, a trend that con-
tinued into the 1800s (Miller 1996; Thondhlana and Martinón-Torres 2009). This period
covers the inception and flourishing of early states, among them Mapungubwe (c. 1220–
1290), Mapela (c. 1000–1400) and Great Zimbabwe (c. 1000–1700) and later ones such as
Torwa-Changamire (c. 1400–1820) and Venda (c. 1700–1900). The boundaries of some of
these states fell within and outside copper production landscapes such as Phalaborwa and
Musina in South Africa (Thondhlana et al. 2016), Dukwe in Botswana (van Waarden
2014), Hurungwe and Copper Queen in Zimbabwe (Swan 2002) and Katanga in
Congo-Kinshasa (Bisson 2000) (see also Chirikure 2018).
Archaeological and archaeometallurgical evidence has shown that various innovations
were associated with copper-based metallurgy over the past 2000 years. For example, while
early first-millennium AD crucibles used at Kansanshi were normal pottery (Bisson 2000),
by the late first millennium and the beginning of the second, custom-made vessels were
used by K2 people near the Umzingwani and Limpopo River Confluence, as well as by
Gumanye populations at Great Zimbabwe (Chirikure et al. 2015; Bandama et al. 2017).
Amidst these innovations, the unchanging observation is that copper and its alloys were
mostly used to make decorative items such as necklaces, bracelets and leglets. However,
the most important question is: what was the value of these objects in the political econ-
omies of some of the states that flourished in southern Africa? In order to address this
question, an archaeometallurgical study was undertaken of remnants of wound copper
and copper alloy artefacts excavated from the sites of Jahunda and Little Mapela in south-
western Zimbabwe. The technological information obtained was used to initiate a conver-
sation around the multiple uses and values embedded in this category of artefacts. The
conclusion reached is that while the flexible copper wire objects were used for decorative
purposes, and may sometimes have been used as status symbols, the production of their
raw materials and the circulation and consumption of the finished objects speaks to wider
social, economic and political networks in which goods, ideas and perhaps people moved
across contiguous and distant landscapes.

Archaeological background: Jahunda and Little Mapela


According to Chimhundu (1992), and unlike some empires in other parts of the world
such as Eurasia, political formations such as states and local variations thereof were not
as expansive in southern Africa as previously imagined. Indeed, research has shown the
existence of multiple and chronologically overlapping social formations in regions such
as southwestern Zimbabwe and its adjacent territories (Beach 1994; Chirikure et al.
2012). While not much is known about most early social formations, the assumption in
southern Africa is that most drystone walled sites were either élite centres or former capi-
tals of extant states and/or chiefdoms (Garlake 1970, 1973; Huffman 2007; Pikirayi 2013).
Using this logic, the drystone walled sites of Jahunda and Little Mapela are likely to have
been centres of power. Oral traditions suggest that Jahunda is a former élite centre of the
ruling Kalanga groups, specifically the Jahunda lineage of Gwanda, which is itself a
4 F. BANDAMA ET AL.

corruption of the name Jahunda. Little Mapela is similarly associated with the ruling
lineages of Kalanga populations that occupied the Shashi region 60 km to the south.
Archaeological excavations were undertaken at these two sites to generate comparative
insights on the materialisation of power and statehood in precolonial southern Africa
(Chirikure et al. 2013, 2014).
The drystone walled (freestanding walls) site of Jahunda (Figure 2) is located near the
modern town of Gwanda in southwestern Zimbabwe. Built in neatly shaped blocks of

Figure 2. Jahunda: site plan showing the location of the excavation trenches.
AZANIA: ARCHAEOLOGICAL RESEARCH IN AFRICA 5

granite, the typology of the walls at Jahunda conforms to the Q-style of Garlake (1970),
which is comprised of neatly coursed drystone masonry built with regular shaped
blocks. Interestingly, this site has a conical tower, albeit smaller than that at Great Zim-
babwe (Manyanga 2007; Pikirayi and Chirikure 2011: 228). Excavations on three
middens at the site recovered a wide range of material, including pottery, bone, glass
beads and an array of flexible copper or copper alloy wire objects. Radiocarbon dates
for charcoal obtained from Trenches 1 and 3 estimate the occupation of this site to
have fallen between cal. AD 1220 and 1430 (Table 1).
The archaeological site of Little Mapela (Fig 3) is located about 60 km south of Jahunda
(Figure 1). Little Mapela has both retaining and free-standing walls, with a thirteenth- to
fifteenth-century occupation. Its small summit occupation is terraced from carefully
selected and coded blocks that end with herringbone decoration at the entrance. It has
freestanding walls on the western side that form enclosures where houses were built.
Cattle kraals are located on the flats situated on the southwestern side of the site. The
site was first excavated by Garlake (1968). Follow up work conducted in 2014 (Figure
3). A series of radiocarbon dates was obtained from charcoal recovered from the exca-
vations (Table 1). Their calibration in Oxcal using the southern hemisphere curve
(SHCal13) (Hogg et al. 2013) produces a sequence spanning the period between cal.
AD 1280 and cal. AD 1460. Excavations at Little Mapela yielded Leopard’s Kopje
pottery, flexible copper/copper alloy wires, crucibles, bone and glass beads.

Materials and methods


A sample of 28 wound wire bangle fragments, two crucible fragments and a crucible slag
specimen (from Little Mapela) were studied using standard archaeometallurgical tech-
niques. Twelve of the wound wire bangle fragments came from Jahunda (Figure 4) and
16 from Little Mapela (Figure 5). The coiled wires were mostly wound on vegetal cores
that are still present in some specimens, having been preserved by the copper. The
wires were a mix of rectangular and rounded types. The crucibles from Little Mapela
are too fragmentary to be reconstructed to their original morphology. Nonetheless, they

Table 1. Radiocarbon dates from Jahunda and Little Mapela calibrated using OxCal 4.3 and the SHCal13
calibration curve (Hoggs et al. 2013). All the dates were run on charcoal.
Site Context Laboratory number Uncalibrated date BP Calibrated date cal. AD
Jahunda Trench 1 Level 2 Beta-326779 600 ± 30 1385-1431 67.7%
1319-1352 27.7%
Jahunda Trench 3 Level 3 Beta-326780 800 ± 30 1220-1290 95.4%

Little Mapela Trench 1 Block C1 Beta-392081 470 ± 30 1597-1611 3.4%


Layer 3 1423-1500 92.0%
Little Mapela Trench 1 Block C1 Beta-392084 490 ± 30 1413-1482 95.4%
Layer 2
Little Mapela Trench 1 Block C1 Beta-392078 540 ± 30 1400-1450 95.4%
Layer 4
Little Mapela Test Pit 3 Layer 1 Beta-392079 620 ± 30 1378-1424 45.5%
1310-1360 49.9%
Little Mapela Test Pit 3 Layer 4 Beta-392082 670 ± 30 1296-1396 95.4%
Little Mapela Test Pit 3 Layer 3 Beta-392077 710 ± 30 1341-1390 48.9%
1280-1326 46.5%
6 F. BANDAMA ET AL.

Figure 3. Little Mapela: site and excavation plan.

conform to several non-specialised examples known throughout the region that resemble
ordinary ceramics with slag attached to the inner walls and vitrification on the outer side
(Miller and Hall 2008; Chirikure et al. 2015; Thondhlana et al. 2016; Bandama et al. 2017).
The crucible slag specimen is vesicular and porous.
After obtaining relevant permits, samples were cleaned, cut, mounted as polished
blocks, ground and polished up to ¼ micron diamond finish following standard metallo-
graphic procedures in the Archaeological Materials Laboratory at the University of Cape
Town. To bring out the microstructure (such as grain boundaries and sizes) the coiled
wires were etched using a ferric chloride etchant for copper-based objects (Scott 1991).
The polished blocks were studied using reflected plane polarised light microscopy using
an Olympus BX51 microscope at varying magnifications from 100 x up to 1000 x magnifi-
cation. Subsequently, samples were then carbon coated to allow the analytical surface to
conduct electrons and to stop specimens from charging during SEM analyses. A LEO
Stereoscan S440 Scanning Electron Microscope (SEM) fitted with a Fison Kenex wave-
length Dispersive X-Ray system housed in the Electron Microscopy Unit at UCT was
used to perform the work. The analyses were carried out in the backscatter mode under
the following conditions: a working distance of 25.0 mm, an excitation voltage of 20.0
kv, a beam current of 1.0 picoAmps, a tilt angle of 0.0 degrees and a takeoff angle of
AZANIA: ARCHAEOLOGICAL RESEARCH IN AFRICA 7

Figure 4. Jahunda: photographs and stereo-micrographs of analaysed specimens.

35.0 degrees. SEM–EDS analyses of the slag and crucibles were performed on areas
measuring ∼70–100 by 100–150 μm in order to obtain fairly representative compositional
data. To obtain compositional data, SEM-EDS measurements of 5–10 analyses per speci-
men were performed and all the results were reported as normalised weight percentages.

Analytical results
Under the stereozoom microscope, surface striation marks of the rounded (circular
section) wires suggested that they were progressively drawn by pulling wire through per-
forated iron draw-plate in cycles that involved reheating (Friede 1975: 234). The rectangu-
lar section wires were mostly made by hammering, the blows of which are identifiable at
high magnification on a stereo-microscope. As far as the results of optical microscopy are
8 F. BANDAMA ET AL.

Figure 5. Little Mapela: photographs of the analysed cuprous samples.

concerned, the etched microstructures of the coiled wires consist of deformed and recrys-
tallised grains with annealing twins (Fig 6) (Oudbashi and Hasanpour 2016). The variable
grain sizes for both the drawn and the hammered wires imply cycles of hot and cold
working. Crucible samples LMA04 and LMA05 were fired from inside (above) because
of the heavy slagging in the interior and microstructurally exhibited instances of magnetite
in the glass matrix.
Turning to chemical analyses, the initial screening using a Niton XLT Portable XRF
machine identified copper and bronze objects as shown in Table 2. Unalloyed copper
dominates the wires from Little Mapela, with only one specimen (LMA03) having
traces of tin. One specimen from Jahunda (JAH06) also produced less than 1% Sn.
Unequivocal bronzes, typically having from 3% and above in tin content (Oudbashi
et al. 2012) were picked by the P-XRF analysis and are ubiquitious at Jahunda where
AZANIA: ARCHAEOLOGICAL RESEARCH IN AFRICA 9

Figure 6. Jahunda (JAH) and Little Mapela (LMA): etched optical microscope micrographs of bronze
(JAH09 and JAH12) and copper (LMA16) coiled wires and unetched SEM backscatter images of crucibles
(LMA04 and LMA05) and crucible slag (LMA06).

the average tin percentage is 7.8, with two specimens (JAH04 and JAH07) suggesting unal-
loyed copper. The tentative nature of these analyses prompted detailed studies in the SEM.
The P-XRF analyses are very much comparable to the SEM-EDS readings from the
same specimens (Table 3). Copper dominates at Little Mapela. Samples for LMA03 and
JAH06 were shown to be low tin bronzes with tin levels of 4.1% and 2.6% respectively.
This shows that the initial lower readings recorded in P-XRF may have been affected by
the surface corrosion. Variable amounts of iron were detected in some of the samples
using the SEM. Iron can be present in raw copper or finished objects in different states
— for example, slag inclusions, oxides, sulphides or metal — and caution should therefore
10 F. BANDAMA ET AL.

Table 2. Normalised P-XRF readings (percentages) for wires from Little Mapela (LMA) and Jahunda
(JAH).
Sample Number Fe Cu Zn Sn Total Metal/alloy
LMA01 2.8 97.2 0.0 0.0 100 Copper
LMA02 5.9 94.1 0.0 0.0 100 Copper
LMA03 4.0 94.2 0.0 0.8 100 Bronze?
LMA07 8.2 91.8 0.0 0.0 100 Copper
LMA08 7.5 92.5 0.0 0.0 100 Copper
LMA09 5.6 94.4 0.0 0.0 100 Copper
LMA10 4.6 95.4 0.0 0.0 100 Copper
LMA11 4.8 95.2 0.0 0.0 100 Copper
LMA12 6.0 94.0 0.0 0.0 100 Copper
LMA13 3.9 96.1 0.0 0.0 100 Copper
LMA14 6.5 93.5 0.0 0.0 100 Copper
LMA15 8.4 91.6 0.0 0.0 100 Copper
LMA16 5.0 95.0 0.0 0.0 100 Copper
JAH01 0.0 93.0 0.0 7.0 100 Bronze
JAH02 0.0 94.0 0.0 6.0 100 Bronze
JAH03 0.0 94.0 0.0 6.0 100 Bronze
JAH04 1.5 98.5 0.0 0.0 100 Copper
JAH05 0.4 88.0 0.0 11.6 100 Bronze
JAH06 1.0 98.1 0.0 0.9 100 Bronze?
JAH07 0.0 100 0.0 0.0 100 Copper
JAH08 0.8 89.2 0.0 10.0 100 Bronze
JAH09 0.0 92.0 0.0 8.0 100 Bronze
JAH10 0.7 92.3 0.0 7.0 100 Bronze
JAH11 0.0 93.0 0.0 7.0 100 Bronze
JAH12 0.5 91.5 0.0 8.0 100 Bronze

Table 3. Jahunda and Little Mapela: SEM-EDS readings for metallic objects and prills trapped in
crucibles and slag.
Material Sample # Fe Cu Zn Sn Total Metal/alloy
Coiled wire fragment JAH01 0.30 91.60 0.00 8.10 100 Bronze
Coiled wire fragment JAH02 0.09 93.11 0.00 6.80 100 Bronze
Coiled wire fragment JAH03 0.80 92.20 0.00 7.00 100 Bronze
Coiled wire fragment JAH04 1.22 98.78 0.00 0.00 100 Copper
Coiled wire fragment JAH05 0.30 86.20 0.00 13.50 100 Bronze
Coiled wire fragment JAH06 0.10 97.30 0.00 2.60 100 Bronze
Coiled wire fragment JAH07 0.90 99.10 0.00 0.00 100 Copper
Coiled wire fragment JAH08 0.00 89.50 0.00 10.50 100 Bronze
Coiled wire fragment JAH09 0.72 91.08 0.00 8.20 100 Bronze
Coiled wire fragment JAH10 0.45 91.05 0.00 8.50 100 Bronze
Coiled wire fragment JAH11 0.23 92.07 0.00 7.70 100 Bronze
Coiled wire fragment JAH12 1.20 90.00 0.00 8.80 100 Bronze
Coiled wire fragment LMA01 0.23 99.77 0.00 0.00 100 Copper
Coiled wire fragment LMA02 0.00 100.00 0.00 0.00 100 Copper
Coiled wire fragment LMA03 0.25 95.70 0.00 4.05 100 Bronze
Prill in crucible fragment LMA04a 1.46 98.54 0.00 0.00 100 Copper
Prill in crucible fragment LMA05a 0.30 96.30 3.40 0.00 100 Brass
Prill in crucible slag LMA06a 0.00 100.00 0.00 0.00 100 Copper
Coiled wire fragment LMA07 0.92 99.08 0.00 0.00 100 Copper
Coiled wire fragment LMA08 0.00 100.00 0.00 0.00 100 Copper
Coiled wire fragment LMA09 2.04 97.96 0.00 0.00 100 Copper
Coiled wire fragment LMA10 0.46 99.54 0.00 0.00 100 Copper
Coiled wire fragment LMA11 2.70 97.30 0.00 0.00 100 Copper
Coiled wire fragment LMA12 1.28 98.72 0.00 0.00 100 Copper
Coiled wire fragment LMA13 2.00 98.00 0.00 0.00 100 Copper
Coiled wire fragment LMA14 0.00 100.00 0.00 0.00 100 Copper
Coiled wire fragment LMA15 1.99 98.01 0.00 0.00 100 Copper
Coiled wire fragment LMA16 0.70 99.30 0.00 0.00 100 Copper
AZANIA: ARCHAEOLOGICAL RESEARCH IN AFRICA 11

be exercised when reporting or interpreting bulk iron levels in ancient copper (Tylecote
et al. 1977: 328).
SEM-EDS analyses also confirmed the presence of standard tin bronzes at Jahunda,
with a tin content averaging 8.2%. This places these objects within the known range of
publised bronzes in southern Africa (Miller 2002, 2003; Bandama 2013; Bandama et al.
2015, 2016, 2017). Two specimens (JAH05 and JAH08) have a tin content suggesting the
production of high-tin bronzes (>10% Sn). Interestingly, although the prills in one of
Little Mapela’s crucible fragments (LMA04a in Table 3) were of pure copper, the
slagged component of this crucible (LMA4b) has enough tin (4.5%) to suggest bronze
working. The same site also intimates brass working (melting or re-melting). This is
captured by one of the prills in the crucible fragment LMA05a that has 3.4% zinc.
Brass is the least known, least represented and perhaps most challenging precolonial
cuprous metallurgy in the region. Its production would have required specialised cruci-
bles (see Roodt 1993; Maggs and Miller 1995) because of its volatility, but open crucibles
were used to melt the alloy at Great Zimbabwe (sample GE05 from the Great Enclosure)
(Bandama et al. 2017).

Discussion: pure copper, bronze, or brass, does it matter?


There is no doubt that copper and copper alloy producers and users in precolonial
southern Africa were aware of the physical and chemical properties of the metals that
they worked and employed. Copper is a soft, ductile, malleable, recyclable, non-magnetic,
corrosion resistant, antibacterial, red metal that turns into shades of green when oxidised.
It can be easily alloyed with tin and zinc to produce bronze and brass. These alloys are, in
turn, harder, stronger and tougher than pure copper, but do retain most of copper’s advan-
tages. Depending on the specific composition of the alloy and how it has been processed,
bronze is golden in colour and highly ductile with low friction and some resistance to cor-
rosion. Its alloying metal (tin) is durable, malleable and non-toxic and has a high resist-
ance to chemical attack, although it is mechanically weak in its pure form (Nattrass
1987: 4). Brass has a reddish-gold colour with a higher malleability than either bronze
or copper. It too is resistant to corrosion with a low friction. Generally, bronze and
brass are much easier to cast than copper alone because they melt at much lower temp-
eratures than copper and generate fewer gasses capable of causing blowholes and porosity
in the finished products (Oudbashi et al. 2012). Technological analyses of flexible wires
from Jahunda and Little Mapela identified objects made of pure copper, tin bronze and
brass. Limited as the analyses are, they nevertheless indicate a dominance of tin bronze
metallurgy at Jahunda in sharp contrast to Little Mapela where there is more evidence
of pure copper mixed with limited amounts of bronze and brass metallurgy. This raises
significant questions about the values of pure copper, bronze and brass and their associ-
ated colours, which are known to be cultural meaningful (Herbert 1984; Chirikure 2015).
We return to this topic below.
Microstructural (OM) and compositional (P-XRF and SEM-EDS) analytical results are
consistent with indigenous technological processes for cuprous metallurgy. Evidence
suggests a combination of cycles of hot and cold working. Most of the evidence,
however, points to hammering and wire drawing. Wire drawing was introduced to
southern Africa towards the end of the first millennium AD (Herbert 1973). It is possible
12 F. BANDAMA ET AL.

that some wires were initially drawn to a certain diameter and hammered to produce the
rectangular cross-section that is characteristic of some of the objects under study. It is not
clear which bronze-making process was used to alloy the tin and copper, but the absence of
unreacted tin or copper oxide inclusions points to the melting of metallic tin and copper
(Tylecote et al. 1989). Presumably, brass was melted in similar crucibles and cast or ham-
mered to produce different objects. The volatility of zinc often forced brass makers to use
constricted (sometimes lidded) vessels (Rehren 2003: 209; Martinón-Torres and Rehren
2014: 115). However, this technological solution mostly works with primary production
and applies less to remelting ready made brass.
What motivated different choices for the dominant use of tin bronzes at Jahunda and
pure unalloyed copper at Little Mapela? This certainly has little to do with access since
Jahunda is 60 km inland when compared to Little Mapela. Superficially, it might appear
as if Jahunda was an élite Zimbabwe tradition centre built in Q-style with the consequence
that it could source bronze, which is traditionally assumed to have been an élite-associated
alloy. However, Little Mapela also has impressive drystone walling decorated with sand-
stone blocks that alternate with grey basalt rocks, thereby creating a stunning impression
from a distance. These differences speak to cultural logics and to the preferences and
values that the inhabitants of the two sites attached to pure copper and tin bronzes.
The redness of copper appealed to one of the important spiritual and gendered colours
in Bantu Africa (Ellert 1984: 5; Herbert 1984; Bvocho 2005: 419; Huffman 2007: 85–87;
Chirikure 2015). The combination of ‘white’ tin with red copper produced a colour that
was more locally valued than gold. That different amounts of alloy components creates
changeable colours and texture, when objects are hammered, heated, polished or
exposed to natural oxidation may have enhanced the appeal of copper and its alloys
(Herbert 1973). While Jahunda is within the gold belt, it appears as if its inhabitants
had no appetite for this metal. This may not be surprising given that copper was much
more deeply entrenched into the local value systen than gold, a metal introduced
through influences from the Indian Ocean trade network.
What was the use and value of flexible wire-wound ornaments in communities resident
at Jahunda and Little Mapela? While we may speculate on their potential archaeological
uses, ethnographic data offer some possible indicators. Ethnographically, ornaments
made of flexible wires satisfied a much wider range of socio-cultural needs that included:
(1) a general purpose currency/store of value; (2) decoration; (3) sexual enticement; (4)
gender/age/status makers; and (5) medicinal/antibacterial uses. Complete flexible wire
objects were exchanged between various groups, thereby assuming the form of a ‘token’
through which value was measured and exchanged. Traditionally, huge amounts of
copper bangles were viewed as a store of wealth. For example, one woman is reported
to have been wearing up to 25 kilograms of expressive metals, an unmistakeable form
of hoarding wealth that cemented the role of metallurgy as a convenient store of value
(Sundström 1974: 227). However, the same applied to owning large heads of cattle,
nuggets of salt and other commodities that could be converted into other services and
goods as and when the need arose. It is important to note that in any exchange relationship
the parties involved had differing expectations of the same transaction (Berry 2007).
Amongst some Shona communities, cuprous bangles were used as payment for minor ill-
nesses to healers (Frommer 2002). The valuation of flexible wire ornaments, like that of
AZANIA: ARCHAEOLOGICAL RESEARCH IN AFRICA 13

the objects for which they were exchanged, was thus situational and context dependent
(Guyer 2007).
A survey of the ethnographic literature from soutrh-central and southern Africa
suggests that the use of flexible wire objects as status indicators was variable. Kings, com-
moners and people of varying social status all had access to these objects. Lunda, Yeke,
Luvale and Shona nobles and their relatives were often observed laden with bundles of
these flexible wires (Bvocho 2005: 410; Musambachime 2017). Archaeologically, one indi-
vidual of high status buried at Danamombe in southwestern Zimbabwe had a large quan-
tity of bangles covering the area from just under the knee to the ankle (Chirikure 2015).
However, amongst the Yeke of Katanga even slaves wore lavish quantities of these wound
wires. In 1867 David Livingstone met a procession of Tippu Tib’s female slaves who had ‘a
jaunty walk, and never gave in on the longest march; many pounds’ weight of fine copper
leglets above the ankles seemed only to help the sway of their walk’ (Waller 1875: 143).
Sometimes, it was the status of the individual that conferred value on the objects (Chiri-
kure et al. 2016), cautioning us against the assumption that all objects have intrinsic value.
Some ethnographic reports suggests that the lavish wearing of cuprous wires produced
audible tinkling and visual twinkling that heralded an unmistakable presence. The tonality
of these sounds was very important in intimacy between sexual partners. An explicit but
figurative example comes from a Shona figure of speech which says ‘Chiripo chiripo,
ndarira (nyamugwe) imwe hairiri,’ which translates to ‘What is there is there, one brass
wire does not tinkle on its own’ (Fortune 1975: 49). Together with other ‘female sexual
ornamentations’ such as glass beads, the tinkling and twinkling of the correct amount
of these wound wires could entice sexual partners (Bvocho 2005: 420). The Tsonga
people of southeastern southern Africa even had specific names (xidzingidzingi, xihati-
mana, xipoyilana, xitorofantini and xikhalamazu) and functions for their version of
these bangles, known as Vusenga (Rikhotso 1985: 185). In a pattern that permeates
many southern African groups, vusenga (ubusenge for Nguni-speakers, liseka for the
Luvale, minungu for the Lunda, insambo for the Bemba, tsambo/ndarira/nyamugwe for
the Shona and vhukunda for the Venda) were mostly used by married women and
mature girls who were about to get married (Rikhotso 1985; Bvocho 2005; Walker
2016; Musambachime 2017). Just like many other ornaments, this category of objects
was an integral part of a multi-layered communication system (Dubin 1987: 119;
Bvocho 2005) in various contexts. However, the anti-bacterial qualities of copper meant
that it was also used for medicinal purposes. Other ethnographic examples show high
ranking officials, as well as slaves, wearing voluminous amounts of wound-wire anklelets,
leglets, necklets, armlets and bracelets, so much, indeed, that their walk was straddled by
the quantity and wieght of the objects (Walker 2016: 80–83; Musambachime 2017).
Archaeometallurgical analyses have demonstrated the presence of copper, bronze and
— to a lesser extent — brass metallurgy in southwestern Zimbabwe. Interestingly, geologi-
cal maps indicate an absence of sources of copper, tin and zinc near Jahunda and Little
Mapela. No substantial evidence of primary copper production evidence is known from
these sites or their immediate surrounds. However, caution is required because not
much research has been done in this area and more archaeometallurgical studies are
required as some slags that might look like waste materials from iron smelting may actu-
ally have derived from copper metallurgy (Miller and Killick 2004). However, in discussing
territories in precolonial Africa, we must be alert to the fact that modern political
14 F. BANDAMA ET AL.

boundaries did not exist in the past. This calls for a consideration of the broader land-
scapes from which populations may have extracted resources and the embedded networks
for the production and circulation of commodities.
Although no copper deposits are known from the immediate surrounds of the sites
understudy, a few sources do occur in greenstone belts (to the north and northwest)
and shear zones (to the south) of the Zimbabwe Craton. Across the border in Botswana,
van Waarden (2014: 1) has reported substantial copper deposits related to the metamor-
phosed volcanic and sedimentary ‘greenstone’ rocks that were mined precolonially at
places such as the Selkirk Mine and Tati Greenstone Belt mines. Much closer to Little
Mapela, she also noted the occurrence of copper deposits in southeastern Botswana’s
Magogaphate Shear Zone (van Waarden 2014: 1). Closer to the town of Gwanda just
north of Jahunda, copper deposits also occur (Map of Mineral Resources of Zimbabwe
1988), but their precolonial exploitation is not yet known. Much further afield, copper
ores occur in the Limpopo Basin at Musina and further south in Phalaborwa, both in
South Africa.
While lead isotope analyses have potential to illuminate the movement of metal across
the landscape, and in the process point to networks of interaction in the past, the domi-
nance of a style of objects in wider area may points to the same. The area roughly 200 km
in radius from where the borders of Botswana, South Africa and Zimbabwe meet is occu-
pied today by Venda, Kalanga, Sotho-Tswana and Tsonga speakers. Interestingly, the
musuku copper ingots that are historically identified with the Venda were recovered in
southwestern Zimbabwe in the areas of Mberengwa and Beitbridge, as well as at
Musina in what is now South Africa. Mberengwa is adjacent to Gwanda. It appears as
if there existed a network through which ideas or finished objects, in this case ingots,
were circulated between various peoples. Such ingots would have been melted to
produce copper objects recovered from Jahunda and Little Mapela. Their circulation
was most likely multi-directional such that more copper was sourced from what is now
modern day Botswana, raising the potential for mixing and remixing of metal from
various areas.
With regard to tin, the precolonial mining of this metal is best known at the unequi-
vocally large-scale mines of Rooiberg in the southern Waterberg Mountains of
Limpopo Province, South Africa (Miller and Hall 2008,; Bandama et al. 2013, 2015).
On current evidence, Rooiberg is the only place in southern Africa from which tin has
been studied and finger-printed to various bronzes in the region (Molofsky et al. 2014).
Little Mapela and Jahunda are within 200 km of Rooiberg. Considering distances in the
past, and the fact that modern political borders did not exist, the inhabitants of southwes-
tern Zimbabwe could easily have accessed this tin through both direct and indirect means.
However, other local sources within the region such as Kamativi, Bikita and Rusape are
also potential candidates for these communities were not closed (Barnes Pope 1938;
Caton-Thompson 1931). The presence of glass beads from the Indian Ocean points to
these long distance networks of down-the-line exchange or direct contact and interaction.
Brass was most likely brought to Little Mapela in this way. This fulfils the expectation that
wherever people meet they exchange ideas, objects, gifts and much more. However, lead
isotope studies of copper and bronze are essential to assist with identifying source geolo-
gies, as a step towards imagining multi-scalar networks in the past.
AZANIA: ARCHAEOLOGICAL RESEARCH IN AFRICA 15

Conclusion
The southwestern Zimbabwean Iron Age sites of Jahunda and Little Mapela yielded
copper-based decorative wirework that was worked using established indigenous tech-
niques. Wires of copper and bronze were drawn and hammered and coiled around
vegetal cores to produce bracelets, necklaces and other related objects. Ores of tin and
copper are not immediately available in areas surrounding the sites. This implies short
and long distance networks of circulation. However, when available, the decorative
objects were used in aesthetic spheres and were important in healing. In some cases,
they were exchanged for other commodities. Therefore, like other aspects of material
culture, the production and circulation of copper-based artefacts followed that of other
objects in broader society. This means that they changed acquired values, lost value and
perhaps regained values depending on context. Consequently, approaching these objects
using broad-based lenses speaks to wider ideas within society and in the political economy.

Acknowledgements
The National Museums and Monuments of Zimbabwe awarded permits for the research. We are
very grateful to the communities in Gwanda and Khami for hosting us over the course of our
research.

Funding
This research was funded by the National Research Foundation of South Africa and the Norway-
South Africa Cooperation through a joint grant to S Chirikure and P D Fredriksen (Grant number:
91317). Additional funds were sourced from the National Research Foundation (Grants: 90524,
91340 and 105 861). The University of Cape Town Research Office also contributed through a
Decoloniality Grant to Chirikure.

Notes on contributors
Foreman Bandama is a Lecturer at Sol Plaatje University and a Post-Doctoral Researcher at the
University of Cape Town, South Africa. He is an African archaeologist with various specialisations
in materials analyses, focusing on metals, glass beads and lithic artefacts. His current research focus
includes the anthropology of technology, the mining and metallurgy of recent (Iron Age) agricul-
turalist communities and precolonial glass beads materials analyses in southern Africa.
Munyaradzi Manyanga is Senior Lecturer in Archaeology and Heritage Management in the
History Department, University of Zimbabwe. He holds a licentiate and PhD from Uppsala Uni-
versity in Sweden. His research interests include adaptation strategies and resilience of Iron Age
and contemporary communities in southern Africa, archaeological theory, landscape archaeology
and community engagement. He is the author of Resilient Landscapes: Socio-Environmental
Dynamics in the Shashi-Limpopo Basin, Southern Zimbabwe c. AD 800 to the Present (Uppsala,
2007) and co-editor with Shadreck Chirikure of Archives, Objects, Places and Landscapes: Multidis-
ciplinary Approaches to Decolonised Zimbabwean Pasts (Langa RPCIG, 2017).
Shadreck Chirikure is Professor of Archaeology in the Department of Archaeology at the University
of Cape Town where he directs the Archaeological Materials Laboratory. His research explores the
political economies of state and non-state formations that existed in precolonial Africa over the last
2000 years. It engages with several themes that include technologies, interaction, trade and
exchange, inequality and among others food ecosystems. In so doing, his work draws on techniques
from the hard sciences and fuses them with interpetations provided by anthropological theories and
16 F. BANDAMA ET AL.

African-centred positions. He is the author of Metals in Past Societies (Springer, 2015) and co-editor
with Webber Ndoro and Janette Deacon of Managing Africa’s Heritage: Who Cares? (Routledge,
2016).

ORCID
Foreman Bandama http://orcid.org/0000-0002-4664-3410

References
Barnes Pope, H. 1938. “The ancient remains of the Makoni District as related to the Zimbabwe
culture.” Proceedings and Transactions of the Rhodesian Scientific Association 36: 91–94.
Bandama, F. 2013. “The archaeology and technology of metal production in the Late Iron
Age of the Southern Waterberg, Limpopo Province, South Africa.” PhD diss., University of
Cape Town.
Bandama, F., Chirikure, S. and Hall, S.L. 2013. “Ores sources, smelters and archaeometallurgy:
exploring Iron Age metal production in the Southern Waterberg, South Africa.” Journal of
African Archaeology 11: 243–267.
Bandama, F. Hall, S. and Chirikure, S. 2015. “Eiland crucibles and the earliest relative dating for tin
and bronze working in southern Africa.” Journal of Archaeological Science 62: 82–91.
Bandama, F., Moffett, A.J. and Chirikure, S. 2017. “Typological and technological attributes of met-
allurgical crucibles from Great Zimbabwe (1000–1700 CE)’s legacy collections.” Journal of
Archaeological Science: Reports 12: 646–657.
Bandama, F., Moffett, A.J., Thondhlana, T.P. and Chirikure, S. 2016. “The production, distribution
and consumption of metals and alloys at Great Zimbabwe.” Achaeometry 58(S1): 164–181.
Beach, D. 1994. The Shona and Their Neighbours. Oxford: Blackwells.
Berry, S. 2007. “Marginal gains, market values, and history (Jane Guyer’s 2004 Marginal Gains:
Monetary Transactions in Atlantic Africa).” African Studies Review 50: 57–70.
Bisson, M. 2000. “Precolonial copper metallurgy: social context.” In Ancient African Metallurgy:
The Sociocultural Context, edited by M. Bisson, P. de Barros, T.S. Childs, and A.F.C. Holl, 83–
146. Walnut Creek: AltaMira Press.
Bvocho, G. 2005. “Ornaments as social and chronological icons: a case study of southeastern
Zimbabwe.” Journal of Social Archaeology 5: 409–424.
Caton-Thompson, G. 1931. The Zimbabwe Culture: Ruins and Reactions. Oxford: Clarendon Press.
Chimhundu, H., 1992. “Early missionaries and the ethnolinguistic factor during the ‘invention of
tribalism’in Zimbabwe.” Journal of African History 33: 87-109.
Chirikure, S. 2015. Metals in Past Societies: A Global Perspective on Indigenous African Metallurgy.
New York: Springer.
Chirikure, S. 2018. “Early metallurgy and surplus without states in Africa south of the Sahara.”
Tagunden des Landesmuseums fur Vorgeschichte Halle 18: 1–14.
Chirikure, S., Bandama, F., House, M., Moffett, A.J., Mukwende, T. and Pollard, A.M. 2016.
“Decisive evidence for multi-directional evolution of socio-political complexity in southern
Africa.” African Archaeological Review 33: 75–95.
Chirikure, S., Hall, S. and Rehren, T. 2015. “When ceramic sociology meets material science: socio-
logical and technological aspects of crucibles and pottery from Mapungubwe, southern Africa”.
Journal of Anthropological Archaeology, 40: 23–32.
Chirikure, S., Manyanga, M., Pollard, A.M., Bandama, F., Mahachi, G. and Pikirayi, I. 2014.
“Zimbabwe Culture before Mapungubwe: new evidence from Mapela Hill, south-western
Zimbabwe.” PLoS ONE 9(10): e0111224.
Dubin, L.S. 1987. The History of Beads from 30000 BC to the Present. London: Thames and Hudson.
Ellert, H. 1984. The Material Culture of Zimbabwe. Harare: Longman Zimbabwe.
Fortune, G. 1975. “Form and imagery in Shona proverbs.” Zambezia 4(ii): 25–55.
AZANIA: ARCHAEOLOGICAL RESEARCH IN AFRICA 17

Friede, H.M. 1975. “Notes on the composition of pre-European copper and copper alloy artifacts.”
Journal of Southern African Institute of Mining and Metallurgy 75: 185–191.
Frommer, C. 2002. “Bodies, capsules and fetishes: the transfer of control over traditional medicianal
knowledge in Zimbabwe.” Montreal Inter-University Initiative Series 5(97): 1–40.
Garlake, P.S., 1968. “Test excavations at Mapela Hill, near the Shashi River, Rhodesia.” Arnoldia 34
(3): 1–29.
Garlake, P.S., 1970. “Rhodesian ruins — a preliminary assessment of their styles and chronology.”
Journal of African History 11: 495–513.
Garlake, P.S. 1973. Great Zimbabwe. London: Thames and Hudson.
Guyer, J.I. 2007. “Africa has never been “traditional”: so can we make a general case? A response to
the articles.” African Studies Review 50: 183–202.
Herbert, E.W. 1973. “Aspects of the use of copper in pre-colonial West Africa.” Journal of African
History 14: 179–194.
Herbert, E.W. 1984. Red Gold of Africa: Copper in Precolonial History and Culture. Madison:
University of Wisconsin Press.
Hogg, A.G., Hua, Q., Blackwell, P.G., Niu, M., Buck, C.E., Guilderson, T.P. and Zimmerman, S. R.
2013. “SHCal13 Southern Hemisphere calibration, 0–50,000 years cal BP.” Radiocarbon 55:
1165–1176.
Huffman, T.N. 2007 Handbook to the Iron Age: The Archaeology of Pre-Colonial Farming Societies
in Southern Africa. Scottville: University of KwaZulu Natal Press.
Maggs, T. and Miller, D.E. 1995. “Sandstone crucibles from Mhlopeni, Natal: evidence of precolo-
nial brass-working.” Natal Museum Journal of Humanities 7: 1–16.
Manyanga, M. 2007. Resilient Landscapes: Socio-Environmetal Dynamics in the Shashe-Limpopo
Basin, Southern Zimbabwe c. AD 800 to the Present. Uppsala: Societas Archaeologica
Uppsaliensis.
Martinón-Torres, M. and Rehren, T. 2014. “Technical ceramics.” In Archaeometallurgy in Global
Perspective, edited by B.W. Roberts and C.P. Thornton, 107–131. New York: Springer.
Miller, D.E. 1996. The Tsodilo Jewellery: Metal Work from Northern Botswana. Cape Town:
University of Cape Town Press.
Miller, D.E. 2002. “Smelter and smith: Iron Age metal fabrication technology in southern Africa.”
Journal of Archaeological Science 29: 1083–1131.
Miller, D.E. 2003. “Archaeological bronze processing in Botswana.” Proceedings of the Microscopy
Society of Southern Africa 33: 18.
Miller, D.E. and Hall, S.L. 2008. “Rooiberg revisited: the analysis of tin and copper smelting debris.”
Historical Metallurgy 42: 23-38.
Miller, D.E. and Killick, D.J. 2004. “Slag identification at southern African archaeological sites.”
Journal of African Archaeology 2: 23–47.
Miller, D.E. and van der Merwe, N.J. 1994. “Early metal working in sub-Saharan Africa: a review of
recent research.” Journal of African History 35: 1–36.
Mineral Resources of Zimbabwe: Base Metal and Industrial Minerals Deposits 1988. 1:1,000,000.
Harare: Surveyor General.
Molofsky, L.J., Killick, D.J., Ducea, M.N., Macovei, M., Chesley, J.T., Ruiz, J., Thibodeau, A. and
Popescu, G.C. 2014. “A novel approach to lead isotope provenance studies of tin and bronze:
applications to South African, Botswanan and Romanian artifacts.” Journal of Archaeological
Science 50: 440–450.
Musambachime, M.C., 2017. Fire-Eaters: Blacksmiths and the Products of the Forge in Pre-Colonial
Zambia. Lusaka: Xlibris Corporation.
Nattrass, G. 1987. “Tin mining in the Transvaal 1905-1914.” MA diss., University of South Africa.
Oudbashi, O, Emami, S.M. and Davami, P. 2012. “Bronze in archaeology a review of the archaeo-
metallurgy of bronze in ancient Iran.” In Copper Alloys — Early Applications and Current
Performance Enhancing Processes, edited by L. Collini, 153–178. Rijeka: INTECH.
Oudbashi, O. and Hasanpour, A. 2016. “Microscopic study on some Iron Age bronze objects from
Western Iran.” Heritage Science 4 (8): 1-8.
18 F. BANDAMA ET AL.

Pikirayi, I., 2013. “Stone architecture and the development of power in the Zimbabwe tradition AD
1270–1830”. Azania: Archaeological Research in Africa 48: 282–300.
Pikirayi, I. and Chirikure, S. 2011. “Debating Great Zimbabwe.” Azania: Archaeological Research in
Africa 46: 221–231.
Rehren, T. 2003. “Crucibles as reaction vessels in ancient metallurgy.” In Mining and Metal
Production Through the Ages, edited by P.T. Craddock and J. Lang, 147–149. London: British
Museum Press.
Rikhotso, F. 1985. Tolo a nga ha vuyi. Braamfontein: Sasavona Publishers.
Roodt, F. 1993. “’n Rekonstruksie van zoeloe geelkoperbewerking by eMgungundlovu.” MA diss.,
University of Pretoria.
Scott, D.A. 1991. Metallography and Microstructure of Ancient and Historic Metals. Los Angeles:
The Getty Conservation Institute.
Sundström, L. 1974. The Exchange Economy of Pre-Colonial Africa. London: C. Hurst and
Company (Publishers) Ltd.
Swan, L. 2002. “Excavations at Copper Queen Mine, northwestern Zimbabwe.” South African
Archaeological Bulletin 57: 64–79.
Thondhlana, T.P. 2013. “Metalworkers and smelting precincts: technological reconstructions of
second millennium copper production around Phalaborwa, northern Lowveld of South
Africa.” PhD diss., University College London.
Thondhlana, T.P. and Martinón-Torres, M. 2009. “Small size, high value: composition and manu-
facture of second millenium AD copper-based beads from northern Zimbabwe.” Journal of
African Archaeology 7: 79–97.
Thondhlana, T.P., Martinón-Torres, M. and Chirikure, S. 2016. “The archaeometallurgical recon-
struction of early second-millennium AD metal production activities at Shankare Hill, northern
Lowveld, South Africa.” Azania: Archaeological Research in Africa 51: 327–361.
Tylecote, R.F., Ghaznavi, H.A. and Boydell, P.J. 1977. “Partitioning of trace elements between ores,
fluxes, slags and metal during the smelting of copper.” Journal of Archaeological Science 4: 305–
333.
Tylecote, R.F. Photos, E. and Earl, B. 1989. “The composition of tin slags from the South-West of
England.” World Archaeology 20: 434–445.
van Waarden, C. 2014. “Prehistoric copper mining in Botswana.” In Encyclopaedia of the History of
Science, Technology, and Medicine in Non-Western Cultures, edited by H. Selin. doi 10.1007/978-
94-007-3934-5_9871-1.
Waller, H. (ed.). 1875. The Last Journals of David Livingstone in Central Africa from 1865 to His
Death. San Fransisco: R.W. Bliss and Company.
Walker, E.J. 2016. “Iron Age decorative metalwork in southern Africa: an archival study.” PhD diss.,
University of Cape Town.

View publication stats

You might also like