Oxford Handbooks Online: Radiocarbon Dating in Rock Art Research

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Radiocarbon Dating in Rock Art Research

Oxford Handbooks Online


Radiocarbon Dating in Rock Art Research
Fiona Petchey
The Oxford Handbook of the Archaeology and Anthropology of Rock Art
Edited by Bruno David and Ian J. McNiven

Subject: Archaeology, Archaeological Methodology and Techniques


Online Publication Date: Mar 2017 DOI: 10.1093/oxfordhb/9780190607357.013.13

Abstract and Keywords

Radiocarbon dating has had a significant impact on rock art research, but an initial
enthusiasm for this dating method by archaeologists has been replaced by a degree of
scepticism. Radiocarbon dates undertaken directly on rock art or on associated mineral
crusts have often reinforced such scepticism, in part because organic carbon-based
materials are present in small quantities and their composition is of such variable
composition that the technique is stretched to its limits. For the researcher planning to
obtain radiocarbon dates, it is essential to have an understanding of the dating options
available, limitations of the technique, the potential impact of their own bias, and the
value of a dating programme that is fully integrated within a larger project. This chapter
outlines the various materials and methods used to radiocarbon date rock art. It includes
some recent examples and highlights some advances as well as shortfalls in the dating of
rock art.

Keywords: Radiocarbon, oxalate, carbon black, Bayesian chronology.

Introduction
Radiocarbon dating remains the most widely applicable, accurate, and reliable
chronometric dating technique available to archaeologists. Radiocarbon (also called 14C)
dating measures the concentration of naturally occurring radioactive 14C that is formed in
the upper atmosphere. 14C is an unstable, or radioactive, isotope because it contains
extra neutrons in its nucleus and, as a consequence, will eventually return to a more
stable 14N isotope by the emission of a negatively charged beta particle. 14C becomes
incorporated into all living things through the nutrients they ingest, and the level of 14C

Page 1 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

in a plant or animal is therefore maintained in near equilibrium until an organism dies.


The decay rate is considered to be constant, and it is therefore possible to estimate the
time since death by comparing the amount of 14C in a dead organism to modern levels.
The technique can date back to around 60,000 years ago, but this varies depending on
the material, laboratory, and experimental conditions (for a detailed introduction to 14C
use in archaeology, see Taylor & Bar Yosef 2014).

The ability to use 14C dating was out of the reach of most rock art researchers until the
introduction of accelerator mass spectrometer (AMS) dating in the late 1970s. This
technique has enabled samples as small as a grain of rice to be routinely dated, which in
turn results in minimal damage to the artefact or surface sampled while also expanding
the range of materials that could be investigated. Despite this advantage, it was more
than a decade beforeHedges et al. (Hedges, Housley, Law, Perry, & Gowlett 1987) and
van der Merwe et al. (van der Merwe, Sealy, & Yates 1987) first applied14C to rock art.
Since then, advances in the AMS technique have enabled ever-smaller quantities (now
micrograms) of carbon to be dated routinely, as well as previously untested materials.
Such refinements have not been undertaken without considerable effort and associated
uncertainty, and critical questions should be addressed by those wishing to demonstrate
the reliability of their dates. These questions include: What is the source and nature of
the organic material to be 14C-dated?What is the reproducibility of 14C values derived
from a given sample? And, what is the relationship of that sample to the event in
question? Unfortunately, researchers investigating rock art have not always adequately
addressed these issues.

In the study of rock art, the identification of exactly what material is being dated is
difficult because most samples are small, amorphous, and of uncertain chemistry.
Consequently, it can be argued that the most significant recent advances in rock art
dating actually fall outside the realm of 14C, instead focussing on material
characterization (Bonneau, Brock, Higham, Pearce, & Pollard 2011; Livingston, Robinson,
& Armitage 2009; López-Montalvo, Villaverde, Roldán, Murcia, & Badal 2014; Mori et al.
2006; Vazquez, Maier, Parera, Yacobaccio, & Sola 2008). Rather than being driven by the
need to demonstrate 14C reliability, the use of chemical characterization techniques such
as gas chromatography, X-ray fluorescence, and Raman spectroscopy has been fuelled by
an interest in the provenance of raw materials and the technical processes used to
prepare the artworks. The introduction of portable machines with improved detection
limits, especially X-ray fluorescence, has enabled the characterization of material in the
field prior to sampling (Beck et al. 2013; McDonald et al. 2014). This in itself promises
greater control and more sample selection opportunities to those investigating the age of
rock art. It also highlights the importance of participation by specialists in each specific
field of analysis from the earliest stages of sample selection.

Page 2 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

The reproducibility and accuracy of rock art dates can also be problematic. Many of the
preparation methodologies used to remove contaminants are, by necessity, very general,
while others—designed specifically for a particular sample type—are experimental. There
is a large body of literature discussing disputed results (e.g., Combier & Jouve 2012;
Gillespie 1997; Nelson 1993; Pettitt & Bahn 2003; Watchman 1999; Valladas & Clottes
2003), in part because of the limited amount of (published) information on sampling
methodology, location, and sample chemistry pre- and post-treatment, but also because of
failure to demonstrate the ability to replicate results with blind and repeated tests
(Gowlett & Hedges 1986). Although researchers have only relatively recently started to
comprehensively address these issues, it should be stressed that the process for the
majority of rock art samples is no more “non-routine” or “experimental” (Pettit & Pike
2007:37) than other dating endeavours.

Last, the relationship of the sample to the target event is critical. This is entirely the
responsibility of the person collecting and submitting the sample for dating and, by
implication, necessitates the researcher understanding the limitations of the preparation
methodologies applied. Unfortunately, disassociation between the event and date is the
most common form of anomaly between 14C age and expected age, usually because of the
complexity of achieving such a goal when dealing with dynamic and complex contexts
(Taylor & Bar Yosef 2014:132–136). Whether it is “premature to construct grand schemes
or make meaningful generalizations” (Pettitt & Pike 2007:28) is open to debate, but this
attitude may in part explain the slow uptake of Bayesian methodologies that could enable
further refinement of chronologies by rock art researchers.

Relative and Associative Dating


Ultimately, the best way to establish the age of rock art is to utilize a holistic approach
that incorporates all dating evidence, both chronometric and relative (e.g.,
superimposition of art works, stratigraphic sequence, stylistic attributes or diagnostic
subject matter, differential weathering and archaeological associations by excavation)
(Aubert 2012; Bednarik 2002). While this multifaceted approach is commonplace for
archaeological investigations generally, these associations are not always obvious for
rock art (Whitley 2012), and perceived stylistic relationships, while useful, can be
contentious with interpretations limited by our own constructs (Aubert 2012).

Regardless, such evaluation has become a valued component of rock art projects, such as
at Nawarla Gabarnmang—a large sandstone shelter in central-western Arnhem Land,
Australia, where occupation dates from the past few hundred years to back beyond

Page 3 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

45,000 cal BP (David et al. 2011). To provide a robust chronological sequence, Gunn et al.
(Gunn, David, Delannoy, & Katherine 2017) plotted patterns of superimposition for
overlapping designs found on three ceiling panels by cross-correlating common artistic
traits. They were able to combine beeswax radiocarbon dates from the identified micro-
layers of superimposed pigment artworks with additional dating information obtained
from a panel that displayed a painted motif thought to represent a horse, which would
have been unknown on the plateau before AD 1845 (Figure 1). Using these relative and
chronometric dating methods, three periods were established for the artworks: that older
than circa AD 1430; between circa AD 1430 and circaAD1640; and between circa AD 1640
and AD 1930. This work demonstrated that the recent artistic repertoire at Nawarla
Gabarnmang included a number of changes in colour and motif types.

An extension of this dating


program was undertaken
using excavated materials
from sediment at the base
of the rock art and spalls
from the art itself to
provide terminus ante or
post quem ages for the
formation of that art. At
Nawarla Gabarnmang, a
Click to view larger
small fragment of a
Figure 1 Nawarla Gabarnmang horse motif from
Panel D1 painted slab was found at
stratigraphic levels
(from Gunn et al. 2017).
containing mixed 14C dates
between 21,252 cal BP and 36,270 cal BP (David et al. 2013), and a sample of charcoal-rich
ash adhering to the back of the pained rock returned an age of 26,913–28,348 cal BP. To
demonstrate the interrelationship of the art and each stratigraphic level, the Nawarla
Gabarnmang dates were coupled with a geomorphological investigation of the cave
structure. This resulted in a detailed understanding of the relative chronology of roof-fall
events, the evolution of the cave, and past interactions of people within the space as they
modified the environment. On this basis, it was concluded that the painting itself was
made shortly prior to a major roof–fall event and dated to circa 28,000–27,000 cal BP
(David et al. 2013, 2014).

Page 4 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

Direct Dating
Most rock art dating has been undertaken on black organic pigments of which charcoal is
assumed to be the primary ingredient, but a wide range of inorganic pigments can be
found, including red, orange, and brown ochre; iron oxides; and black manganese oxide
or hydroxide. These can be dated if an organic binder is present in the paint matrix. The
interpretation of dates from any material is complicated by the fact that we are not
necessarily dating the painting event, but the uptake of 14C prior to death by the
organism used in the paint. In most cases, the difference between the age of the targeted
event and that of the material dated is small, but large offsets are possible, as was the
case for bird track designs from the Canning Stock Route in the Western Desert of
Australia. These designs were expected to date to the time of early nineteenth- and
twentieth-century droving activity along the route, but ages of 12,970±270 BP,
12,620±460 BP, and 5520±290 BP suggested to McDonald et al. (2014:200) that the paints
used contained a petroleum-based binder resulting in values that reflected this ancient
14C source.

Regardless of the material, when selecting samples for dating, it is essential that the
extraction of organic carbon and subsequent purification is free from contamination. For
rock art samples, this is nearly impossible because of the large surface area, amorphous
nature of the paint, and small sample sizes; the researcher has to consider the potential
inclusion of fungi, algae, lichens and microbes, insects, rootlets, dust, and soil (Ridges,
Davidson, & Tucker 2000), all of which can be found naturally on rock surfaces, while the
very presence of humans will often accelerate the growth of a wide range of organisms
on the applied paint (Valladas 2003). There is also the possibility of subsequent animal
and anthropogenic activity. This can include both ancient (e.g., at Cova Remigia in Spain,
where a two-colour paint combination was due to ancient retouching and produced art
work that deviated from the more typical monochromatic drawings of the region [López-
Montalvo et al. 2014]) and modern activities (e.g., the application of kerosene during
modern times to aid photography at Great Gallery, Horseshoe Canyon, Utah [Chaffee,
Hyman, & Rowe 1994]). Often the impact of recent, chemically diverse contamination
may be complex, and care needs to be taken when evaluating the impact of different
contaminants (e.g., burning of tyres could result in either old or young results depending
on whether they were manufactured from fossil petrochemical sources or natural modern
materials; Steelman, Carrera Ramírez, Fábregas Valcarce, Guilderson, & Rowe [2005:
387] argued that a young age for site M10 at Coto dos Mouros, northwest Iberia, was
caused by contamination from burning tyres made from natural rubber products).

Page 5 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

Protocols for sampling and reporting rock art dates are well documented (e.g., Steelman
& Rowe 2012:572–573; Watchman 1999) and are not repeated here. The primary concern
of these authors has been to ensure minimal damage to the artwork while maximizing
sample size and the retention of valuable information. These works are undoubtedly
useful guides, but the wide variety of datable materials and environmental conditions
encountered mean that they can only be taken as a rough indication of what is required.
Without prior knowledge of a site, the artwork, and potential contaminants, it is very
difficult to sample effectively, as demonstrated by the complexity encountered when
dating artwork from along the Canning Stock Route (McDonald et al. 2014). Here, where
possible, the sampling strategy included the measurement of replicates from a single
motif or superimposed paint layers, as well as rock samples from adjacent unpainted
areas to investigate natural levels of organic contamination. Of the “charcoal” drawings
sampled, there was only a 53% dating success rate, and few of the collected charcoal
samples produced sufficient carbon to enable results to be duplicated adequately or at all
(e.g., duplicate samples from a black snake outline at Pinpi 5, in the Jilakurru Ranges,
weighing 6 milligrams and 8 milligrams, returned results of 225±40 BP and 470±270 BP,
respectively. The large error range on the 8 milligram sample is indicative of extremely
low levels of carbon).

Charcoal and Carbon Black

Charcoal is a very popular material for dating. This stems from the fact that it has a high
carbon content and is relatively chemically stable. Charcoal dating, is not, however,
without issues. Of particular concern when dealing with charcoal, and derivatives such as
soot, is the old wood and/or old charcoal effect (also called inbuilt age) that can add an
unknown degree of older carbon into a sample depending on the age of the plant
combusted. Related to this is a storage effect by which charcoal or other materials may
have been curated for some time before being utilized in the paint (Schiffer 1987). In
both cases, the exact magnitude of error is impossible to predict because the impact
depends on the age of the art and contaminants. An extreme example of this has been put
forward by Combier and Jovue (2012), who suggested that the addition of fossil carbon in
the form of lignite or bitumen caused artificially old (circa 32,000 BP) dates from Chauvet-
Pont-d’Arc Cave in the Ardèche region of southern France. Such a hypothesis can be
controversial and difficult to prove either way, in part because of complexities associated
with small sample size and contaminant removal that can result in disparate duplicate
analyses, but also because of complexities with site interpretation. Indeed, this criticism
has been put forward for the Chauvet Cave chronology despite the fact that it has been
subjected to an intensive dating regime involving multiple dating techniques, test

Page 6 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

excavations, and superimposition studies that give confidence to the chronological


evaluation (Clottes & Geneste 2012; see also Pettitt & Bahn 2015; Quiles et al. 2016).

The pretreatment of charcoal commonly involves a succession of ‘acid-base-acid’ washes


that vary in length and intensity according to sample size and material robusticity. These
pretreatments theoretically remove humic acids (a major organic component of soil made
up of a mixture of different acids formed from the decay of plant and microbial remains),
carbonates (introduced via groundwater contact with limestone and other carbonaceous
sediments), fulvic acids (humic acids of lower molecular weight and higher oxygen
content), and atmospheric carbon dioxide (CO2) that has been absorbed onto the surface
of the sample. A low-temperature (circa 300°C) combustion step may also be added to
remove the most liable fraction (Bird et al. 2010; Sand et al. 2006; Valladas 2003).
Success is, however, not guaranteed: chemical characterization of pretreated samples
has demonstrated that contaminants may remain that cannot always be adequately
accounted for (Alon, Mintz, Cohen, Weiner, & Boaretto 2002; Livingston et al. 2009).

Most black rock art samples submitted for 14C analysis are highly amorphous, burnt
organic materials of unknown origin. Commonly called carbon black, this terminology
hides the fact that this material can originate from a wide variety of sources, including
incomplete combustion products of substances such as wood, charcoal, fruit, bone,
resins, oil, and fat (also variously termed ‘vegetable or plant black’, ‘ivory/bone black’, or
‘lampblack’) (Bonneau et al. 2011). Such materials are less stable chemically than
charcoal and often completely dissolve if routine pretreatment methods are used.
Moreover, contaminants, including carbon-containing oxalate minerals (CaC2O4) derived
from some rock surfaces (see the later section on Mineral Accretions), may become
preferentially concentrated in the sample during pretreatment (Hedges et al. 1998:36–
37).

Many experimental pretreatments have been devised to forgo or minimize these wet
chemistry steps. One such technique is plasma-oxidation ashing. This uses a low-
temperature, low-pressure oxygen plasma that is excited by an oscillating electric field.
The temperature of the gas components is increased by elastic collisions between the
electrons and the gas, which result in the electrons becoming sufficiently energetic to
break molecular bonds while the gas remains at temperatures below 150°C (see Russ,
Hyman, Shafer, & Rowe 1990, 1991 and Rowe 2009 for methodology). Theoretically, this
methodology prevents the decomposition of carbonate and oxalate minerals that occur at
the higher temperatures (circa 800°C) used during more traditional radiocarbon
combustion techniques. This process can be used to remove organic contaminants by the
controlled reduction of a percentage of the starting mass prior to recombustion at a
higher temperature (Bird et al. 2010) or to selectively remove and date tiny amounts of

Page 7 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

organic binder from the surface of a sample (Rowe 2009). Researchers are divided on the
reliability of plasma-oxidation. Steelman and Rowe (2012:577–578) concluded that
replicate measurements on the same image suggested an uncertainty of ±250 years and
that their tests were more successful in limestone environments than in sandstone
environments. Investigations by Bird et al. (2010) also indicated that plasma-oxidation
could not discriminate between similar organic components. This means that the wet
chemistry approaches used to remove humics prior to combustion are still recommended
if possible (cf. McDonald et al. 2014).

One novel approach to removing contaminants was implemented by Bonneau et al. (2011)
who used a density separation technique previously used on shells (Douka, Hedges, &
Higham 2010) to remove calcium oxalates (density of 2.0–2.2 g/cm3) from black pigments
(density 1.5 g/cm3). Although Fourier transform infrared (FTIR) spectroscopy indicated
that the black pigment was free from contaminants, the expense of the heavy liquid
(sodium polytungstate) and the possibility that the calcium carbonates and oxalates had
chemically bonded to the black pigment meant that the technique was not considered to
be viable. Experimentation with different preparation methodologies to find an
uncontaminated form of autochthonous carbon is sure to continue.

Confirmation of minimal post-depositional contamination using chemical characterization


of both the paint and unpainted rock substrateis becoming more common in rock art
dating. Bonneau et al. (2011) used both Raman spectroscopy and FTIR to evaluate the
presence of calcium oxalates in black, noncharcoal pigments from RSATYN2, a painted
rockshelter in the Drakensberg Mountains (South Africa). They evaluated the success of
different pretreatments to ensure the removal of calcium oxalate (Raman peak at 1,474
cm-1) and gypsum (peak at 1,005 cm-1) prior to dating. They were able to confirm by
Raman peaks at ~1,360 cm-1 and ~1,580 cm-1 that the samples were typical of
amorphous carbon, while the absence of characteristic calcium oxalate (3,600–3,400
cm-1, 1,616 cm-1, 1,315 cm-1) and calcium carbonate (1,429 cm-1, 877 cm-1, and 713 cm-1)
peaks in the FTIR spectra after pretreatment indicated the success of the acid treatment
(Figure 2).

Page 8 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

As noted earlier,
improvements in analytical
techniques have seen the
advent of portable systems
that enable chemical
evaluation in the field
prior to sampling. Beck et
al. (2013) tested a portable
X-ray fluorescence (XRF)
Click to view larger elemental analyser and a
Figure 2 Fourier transform infrared (FTIR) spectra Raman spectrometer to
of rock art sample taken from RSATYN2, in the
Drakensberg Mountains showing progressive
investigate art from the
contaminant removal after different acid treatments caves of Villars and
(H2SO4 and HCl) (x: Wavenumber cm-1; y: FTIR Rouffignac in the
intensity)
Dordogne region of
(adapted from Bonneau et al. 2011:424, figure 6).
France. Although XRF
cannot detect carbon
directly, the absence of heavier elements such as manganese or iron—which are black
and therefore difficult to distinguish from carbon—lent support to the carbon nature of
the black pigments and enabled the researchers to isolate and sample scorch marks on
the rock surface that are thought to have originated from torches. Raman spectrometry,
on the other hand, enabled the detection of organic matter and could be used to separate
carbon additions from black inorganic pigments.

Other Inclusions

Inclusions in paints and other dated surfaces (e.g., mineral accretions) are rare, but can
include plant fibres, pollen, spores, insects, and microorganisms. Generally, inclusions
will originate from a heterogeneous range of processes and sources that are unrelated to
the artwork, potentially making date interpretation complex. It is unusual to find
undisputed evidence of deliberate additions such as those found at the Yam Camp site
near Laura, Australia. Here, Watchman and Cole (1993) recovered a quantity of matted
plant fibres mixed with the quartz-kaolinite paint. Typically inclusions are few and careful
evaluation of context prior to sampling is essential to ensure no cross-contamination has
occurred via contact with visitor’s hands and clothes, or animals (Bednarik 2002:12).

Page 9 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

Pigment Binders

The dating of non–charcoal-derived binders, including animal fats, egg whites, blood and
honey, is highly contentious. These materials often do not survive pretreatment, or
indeed environmental conditions, and/or it can be difficult to chemically characterize and
distinguish them from contaminants (Bednarik 2014:230; Gillespie 1997; Livingston et al.
2009). Of note is the oft-reported dating of a dark red pigment at Laurie Creek (Northern
Territory, Australia)—identified as human blood—that returned a 14C age of 20,320
+3100/−2300 BP (Loy et al. 1990). Subsequent tests on the sandstone surface indicated
that datable organic matter was also present in unpainted areas of the shelter (Nelson
1993), while the small sample size and, in hindsight, processing of the protein using
ultrafiltration techniques that are prone to humectant contamination (see Bronk Ramsey,
Higham, Bowles, & Hedges 2004) could also have contributed to the old apparent age
(see Gillespie 1997).

Quantitative analysis of these compounds (e.g., lipids, proteins, or carbohydrates) by gas


or liquid chromatography mass spectrometry provides greater detail when dealing with
complex mixtures of organic materials and has been used to confirm the presence and/or
removal of degradation products prior to dating (e.g., Livingston et al. 2009; Mori et al.
2006). Vázquez et al. (2008) demonstrated the potential of such characterization. They
removed two black (manganese and iron oxide) samples from camelid figures at Alero
Hornillos 2, Argentina. By isolating lipids using chloroform-methanol and subsequent
analysis by gas chromatography mass spectrometry, they were able to identify saturated
fatty acids characteristic of degraded animal fats, most probably from a ruminant animal
source. While no 14C dates were obtained, such works hold promise for future avenues of
14C dating fraction-specific materials.

Beeswax

Surprisingly, beeswax figures are the most commonly dated art media in Australia (~48%
of radiocarbon dates obtained on rock art), despite its limited geographic distribution to
the northern and western regions of the country (Langley & Taçon 2010). Typically,
‘beeswax’ is a variable mixture of plant resins, gums, and wax. The dating of this material
has gained popularity because it is easy to sample and has few interpretative limitations
because the age of the wax will approximate the age of creation by bees, it is produced
seasonally, and use and collection are concurrent since it becomes brittle and unusable
over time. Theoretically, this material is also poorly soluble and is resistant to microbial,
bacterial, and fungal growth. Tests on the same sample (BW-4 from the site of

Page 10 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

Gunbilngmurrung, western Arnhem Land, Australia), but utilizing different preparation


methodologies (HCl/NaOH/HCl [Nelson, Chaloupka, Chippindale, Alderson, & Southon
1995] vs. acidified (KMnO4) permanganate [Watchman & Jones 2002]) gave ages 400
years apart (4040±90 BP and 4460±80 BP, respectively). This discrepancy remains
unsolved but may be related to differences in sample preparation or to the survival
characteristics of the beeswax (Bednarik 2002:11). The antiquity of these dates is also
anomalous in that most beeswax sampled from northern Australia date to less than 2,000
years ago (Taçon et al. 2010:2; Watchman & Jones 2002:146).

A combination of superimposition studies with the dating of beeswax figures has resulted
in a re-evaluation of early European contact events. Taçon et al. (2010) took a total of 10
beeswax samples from the Djulirri art site in the Wellington Range, Arnhem Land,
Australia, including two samples from a human figure covered by a yellow/orange emu
painting; two beeswax pellets over a painting of a European tall ship; one piece of
beeswax from a figure with hands on hips and wearing a hat; two pieces from a snake
that overlays a large yellow painting of a prau (assumed to be of a southeast Asian sailing
vessel); one piece from a female human-like figure over a white painting of a prau; a
sample from a line that was above both the beeswax snake and over the white prau; and a
sample from an unidentifiable design under the tall ship. The results enabled Taçon et al.
(2010) to determine that the painting of the yellow prau was made prior to AD 1664 and
was therefore not only the oldest dated contact rock art depiction from Australia, but also
some of the earliest evidence for Southeast Asian visits to northern Australia. Similarly,
the dates for the beeswax human figure wearing a hat with hands on hips and the
painting of the tall ship indicated that a close encounter between local Aboriginal people
and Europeans probably occurred in the AD 1700s, much earlier than AD 1818 as
presupposed, opening up the possibility of Dutch rather than British contact at this time.

Indirect Stratigraphic Dating


Indirect stratigraphic dating refers to the dating of rock art using materials or objects
from above or below the artwork (e.g., laminates such as oxalate skins and calcite
[CaCO3] crusts or flowstone and insect nests). Such dating, although not directly related
to the painting event, is often used to bracket the age of artworks and is, therefore, often
characterized by broad age estimates that are further complicated by the carbon source
of the dated materials. It does, however, provide information when there are few other
dating options available.

Page 11 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

Mineral Accretions

When dating flowstone, it is important to recognize the possibility of the incorporation of


“dead-carbon” derived from ancient carbonate rocks, plus any “rejuvenation” that may
have occurred: a process by which new carbon can be added to the cave surface through
the dissolution and re-precipitation of calcite associated with changes in wet/dry
conditions, possibly resulting in younger ages or a mixture of ages (Plagnes et al.
2003). In an attempt to account for such an offset in a nested diamond engraving at
Billasurgam Cave in southern India, Taçon et al. (2013) dated what appeared to be a
recently precipitated dry shawl of calcite in a nearby section of the cave. The resultant
900 ± 30 BP age indicated the magnitude of offset due to ancient carbon input. This
correction was applied to dates on flowstone that bracketed the engraving and resulted
in an age that was consistent with similar engravings of Mesolithic age nearby.

Oxalate minerals (whewellite and weddellite) that can form as crusts over artwork or can
contaminate carbon samples, as already discussed, can also provide dating opportunities.
Oxalates are isolated in the laboratory by treatment in warm, acidified permanganate,
which turns the oxalate into CO2 but leaves carbon from other sources such as amino
acids, peptides, and carbonates (Hedges et al. 1998; method reported in Gillespie 1997).
Therefore, this extraction methodology is more selective than plasma-oxidation or
traditional combustion methodologies. Watchman (1987) reported the first dates on
oxalate crusts from rockart sites in Australia. The methodology was subsequently
improved by isolating individual microscopic layers in stratigraphic sequence to provide
additional dating control and limit cross-contamination (Ruiz et al. 2012; Watchman 1993;
Watchman & Campbell 1996).

Most researchers assume that metabolic processes in algae and microorganisms form
oxalate minerals, but this remains speculative. Consequently, the age of carbon in these
accretions may be near contemporaneous with initial formation or could take place over
many years and theoretically could include both rejuvenation and re-deposition of carbon
(Bonneau et al. 2011; Watchman 1993:468; Watchmen & Campbell 196:411). Regardless,
calcium oxalate dating can be a valuable tool for confirming chronological limits for
pictographs that cannot be otherwise dated directly (e.g., the iron oxide painted “eye-
idols” at Abrigo de los Oculados, Spain [Ruiz et al. 2012] or engraved artworks such as at
Yiwarlarlay in the Northern Territory, Australia [Watchman, David, McNiven, & Flood
2000]).

Page 12 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

Insect Nests

Mud-dauber wasp and termite nests have also been used to provide age-delimiting
radiocarbon ages for artwork (Bednarik 2014; Brady, Thorn, McNiven, & Evans 2010).
Both types of insect gather mud that contains pollen and other organics from nearby
environments, and their nests are particularly common in tropical or subtropical
locations. In some instances, silica or oxalate minerals that can also be dated may
mineralize the nests over time. Often, however, in these instances the interval separating
the age of the dating event and the mineralization of the nest may be so great as to be of
limited use. This was the case for an artwork dated by Bednarik (2014:230) at Princess
Charlotte Bay, north Queensland, Australia. Here, a silicified mud wasp nest underlay the
white kaolinite painted lines of artwork documenting early pearling activities towards the
end of the nineteenth century. The resultant 14C age for organics removed from the nest
could only support an age for the artwork sometime in the past 19,000 years!

Date Evaluation: Bayesian Methodologies


The use of Bayesian statistical methodologies for the evaluation of 14C dates and context
is the latest revolution in 14C dating (Bayliss 2009). The use of Bayesian modelling allows
the statistical scatter on the radiocarbon dates to be taken into account, enables the
dating to become integrated with archaeological observations, and provides a useful
model that can be tested rather than relying on often biased visual inspection. When
correctly applied, Bayesian modelling of radiocarbon dates has demonstrably reduced the
timescales under discussion (e.g., Bayliss, Bronk Ramsey, van der Plich, & Whittle 2007).
However, the development of Bayesian models is not easy and requires considerable
effort to be expended on simulation and model-building, thus requiring a programme of
gradual and progressive dating to infill and test various aspects of the model. Moreover,
the radiocarbon dates that constitute these models are tightly linked to pre-existing
archaeological opinions (i.e., “prior beliefs”), so, more than ever before, the onus is on
the archaeologist to ensure that those beliefs are valid and to recognize these models for
what they are—models that require further testing (Bayliss 2009:127). Comprehensive
guidelines for archaeologists selecting samples for Bayesian evaluation are given in
Bayliss (2015).

The progressive dating methodology and rigorous sample selection protocols


recommended by Bayliss (2009, 2015) do not mould well with the small sample sizes and
complex and often ambiguous superimpositions encountered in rock art. Consequently,

Page 13 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

most rock art chronologies are still largely based on graphic interpretation of individual
radiocarbon-calibrated age ranges. The potential value of Bayesian chronologies has been
highlighted by work at the Billasurgam Cave complex, southern India. Here, Taçon et al.
(2013) dated five pieces of flowstone associated with an engraving consisting of three
interlinked concentric diamond patterns (Figure 3A). Samples C1, C2, and C3 overlay the
engraving, C2 was thought to be younger than C1, and C3 older. C4 was taken from
above these, whereas C5 was taken from a surface that was considered to underlie the
painting (though about 1 metre away) (Taçon et al. 2013:1791). To refine the age of the
engraving, the authors then integrated the relative age sequence they had identified into
a Bayesian model using the program OxCal (Bronk Ramsey 2001). In the model, sample
C5 was designated as predating the engraving and C4 as postdating it. The other three
ages were incorporated within a single phase that Taçon et al. thought correlated most
closely to the period when the engraving was created. This sequence produced a
probable age for the engraving of between 5400 and 5000 cal BP, a significant refinement
when compared to visual inspection of the calibrated age ranges of the unmodelled dates
(Figure 3B).1 However, they noted that C2 produced a very low individual agreement
index (A; an index that gives a measure of how well the date fits within the model. The
threshold for a good agreement is 60%), suggesting that the date was too young, most
likely because of rejuvenation and/or contamination.

Page 14 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

Click to view larger

Figure 3 A. Calibrated radiocarbon ages for dated


samples from Billasurgam Cave. B. Bayesian age
model of the results, incorporating the relative
sequence information as discussed in the text. Grey
= original age distribution; Black = age distribution
based on model constraints. TPQ is the acronym for
terminus post quem

(figure adapted from Taçon et al. 2013: Figure 10).

Conclusion
Radiocarbon dating rock art is difficult; it pushes every parameter of the technique,
requiring attention to detail by the archaeologist and the collaborating 14C scientists.
Over the past few decades, a range of processing techniques for rock art dates have been
developed, but there is still no consensus on how to handle the diverse range of materials
incorporated into these art works, with alternative approaches favoured by different
laboratories and researchers. While most studies of rock art recognize the need for
replication of dates or comparison of sample preparation methodologies, the science is
still hampered by sample size and sample chemistry. Advances in chemical
characterization, especially the use of portable technologies and improved analytical
detection, combined with the dating of sample-specific fractions, nevertheless promises
to open up new avenues of 14C enquiry.

The relationship of a radiocarbon date to the artistic event of interest is also often
difficult to interpret because many dates are of materials that either pre-or postdate the
activity in question. While this problem is not specific to rock art dating, the microscopic

Page 15 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

vertical scale of superimpositions often increases the difficulties of interpretation.


Ultimately, researchers must come to terms with the idea that radiocarbon dating does
not provide an absolute solution to the dating question; rather, it is a tool from which we
can develop chronological models in association with other information gathered from the
rock art site.

Acknowledgements
I wish to thank the many collaborators who have given me the opportunity to be involved
in a diverse and interesting array of thought-provoking projects, in particular the Jawoyn
Association and Margaret Katherine, the senior traditional owner of Nawarla
Gabarnmang. Adelphine Bonneau kindly gave permission for the reproduction of Figure
2. I would also like to thank the editors of the Oxford Handbook of the Archaeology and
Anthropology of Rock Art for the opportunity to put forward my opinions and discuss the
many issues that should be considered before radiocarbon dating images. Last, but not
least, I would also like to give special mention to David Petchey who was my mentor in all
things and inspired my interest in archaeology and solving problems.

References
Alon, D., Mintz, G., Cohen, I., Weiner, S., & Boaretto, E. (2002). The use of Raman
spectroscopy to monitor the removal of humic substances from charcoal: Quality control
for 14C dating of charcoal. Radiocarbon, 44, 1–11.

Aubert, M. (2012). A review of rock art dating in the Kimberley, Western Australia.
Journal of Archaeological Science, 39, 573–577.

Bayliss, A. (2009). Rolling out revolution: Using radiocarbon dating in archaeology.


Radiocarbon, 51(1), 123–147.

Bayliss, A. (2015). Quality in Bayesian chronological models in archaeology. World


Archaeology, 47(4), 677–700. doi:10.1080/0043(8243).(2015).1067640

Bayliss, A., Bronk Ramsey, C., van der Plicht, J., & Whittle, A. (2007). Bradshaw and
Bayes: Towards a timetable for the Neolithic. Cambridge Archaeological Journal, 17(1), 1–
28.

Bednarik, R. G. (2002). The dating of rock art: A critique. Journal of Archaeological


Science, 29(11), 1213–1233.

Page 16 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

Bednarik, R. G. (2014). Mud-wasp nests and rock art. Rock Art Research, 31(2), 225–231.

Beck, L., Genty, D., Lahlil, S., Lebon, M., Tereygeol, F., Vignaud, C., … Paillet. P. (2013).
Non-destructive portable analytical techniques for carbon in situ screening before
sampling for dating prehistoric rock paintings. Radiocarbon, 55(2–3), 436–444.

Bird, M. I., Charville-Mort, P. D. J., Ascough, P. L., Wood, R., Higham, T., & Apperley, D.
(2010). Assessment of oxygen plasma ashing as a pre-treatment for radiocarbon dating.
Quaternary Geochronology, 5, 435–442.

Bonneau, A., Brock, F., Higham, T., Pearce, D. G., & Pollard, A. M. (2011). An improved
pretreatment protocol for radiocarbon dating black pigments in San rock art.
Radiocarbon, 53(3), 419–428.

Brady, L. M., Thorn, A., McNiven, I. J., & Evans, T. A. (2010). Rock art conservation and
termite management in Torres Strait, NE Australia. Rock Art Research, 27, 19–34.

Bronk Ramsey, C. (2001). Development of the radiocarbon calibration program OxCal.


Radiocarbon, 43(2A), 355–363.

Bronk Ramsey, C., Higham, T., Bowles, A., & Hedges, R. (2004). Improvements to the
pretreatment of bone at Oxford. Radiocarbon, 46(1), 155–163.

Chaffee, S. D., Hyman, M., & Rowe, M. W. (1994). Vandalism of rock art for enhanced
photography. Studies in Conservation, 39(3), 161–168.

Clottes, J., & Geneste, J.-M. (2012). Twelve years of research in Chauvet Cave:
Methodology and main results. In D. McDonald & P. Veth (Eds.), A companion to rock art.
(pp. 583–604). Chichester, UK: John Wiley & Sons. doi:10.1002/9781118253892.ch33

Combier, J., & Jouve, G. (2012). Chauvet Cave’s art is not Aurignacian: A new
examination of the archaeological evidence and dating procedures. Quartär, 59, 131–152.

David, B., Barker, B., Delannoy, J.-J., Geneste, J.-M., Petchey, F., & Lamb, L. (2014). A
Pleistocene charcoal drawing or painting from northern Australia. INORA, 69, 18–22.

David, B., Barker, B., Petchey, F., Delannoy, J.-J., Geneste, J.-M., Rowe, C., … Whear, R.
(2013). A 28,000 year old excavated painted rock from Nawarla Gabarnmang, northern
Australia. Journal of Archaeological Science, 40, 2493–2501.

David, B., Geneste, J.-M., Whear, R. L., Delannoy, J.-J., Katherine, M., Gunn, R. G., …
Matthews, J. (2011). Nawarla Gabarnmang, a 45,180±910 cal BP site in Jawoyn country,
southwest Arnhem Land plateau. Australian Archaeology, 73, 73–77.

Page 17 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

Douka, K., Hedges, R. E. M., & Higham, T. F. G. (2010). Improved AMS 14C dating of shell
carbonates using high-precision X-ray diffraction (XRD) and a novel density separation
protocol (CarDS). Radiocarbon, 52(2), 735–751.

Gillespie, R. (1997). On human blood, rock art and calcium oxalate: Further studies on
organic carbon content and radiocarbon age of materials relating to Australian rock art.
Antiquity, 71(272), 430–437.

Gowlett, J. A. J., & Hedges, R. E. M. (1986). Lessons of context and contamination in


dating the Upper Palaeolithic. In Gowlett & Hedges (Ed.), Archaeological results from
accelerator dating (pp. 63–71). Monograph 11. Oxford: Oxford University Committee for
Archaeology.

Gunn, R. G., David, B., Delannoy, J.-J., & Katherine, M. (2017, in press). The past 500
years at Nawarla Gabarnmang, central-western Arnhem Land. In B. David, P. Taçon, J.-J.
Delannoy, & J.-M. Geneste (Eds.), The archaeology of rock art in western Arnhem Land,
Australia. Canberra, AUS: ANU Press.

Hedges, R. E. M., Bronk Ramsey, C., van Klinen, G. J., Pettitt, P. B., Nielsen-March, C.,
Etchegoyen, A., Fernandex Biello, J. O., Boschin M. T., Llamazares, A. M. (1998).
Methodological issues in the 14C dating of rock paintings. Radiocarbon, 40(1), 35–44.

Hedges, R. E. M., Housley, R. A., Law, I. A., Perry, C., & Gowlett, J. A. J. (1987).
Radiocarbon dates from the Oxford AMS system: Archaeometry datelist 6. Archaeometry,
29(2), 289–306.

Langley, M. C., & Taçon, P. S. C. (2010). The age of Australian rock art: A review.
Australian Archaeology, 71, 70–73.

Livingston, A., Robinson, E., & Armitage, R. A. (2009). Characterizing the binders in rock
paintings by THM-GC-MS: La Casa de Las Golondrinas, Guatemala, a cautionary tail for
radiocarbon dating. International Journal of Mass Spectrometry, 284, 142–151.

López-Montalvo, E., Villaverde, V., Roldán, C., Murcia, S., & Badal, E. (2014). An
approximation to the study of black pigments in Cova Remigia (Castellón, Spain).
Technical and cultural assessments of the use of carbon-based black pigments in Spanish
Levantine Rock Art. Journal of Archaeological Science, 52, 535–545.

Loy, T., Jones, R., Nelson, D., Meehan, B., Vogel, J., Southon, J., & Cosgrove, R. (1990).
Accelerator radiocarbon dating of human blood proteins in pigments from late
Pleistocene art sites in Australia. Antiquity, 64, 110–116.

Page 18 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

McDonald, J., Steelman, K. L., Veth, P., Mackey, J., Loewen, J., Thurber, C. R., &
Guilderson, T. P. (2014). Results from the first intensive dating program for pigment art
in the Australian arid zone: Insights into recent social complexity. Journal of
Archaeological Science, 46, 195–204.

Mori, F., Ponti, R., Messina, A., Flieger, M., Havlicek, V., & Sinibaldi, M. (2006). Chemical
characterization and AMS radiocarbon dating of the binder of a prehistoric rock
pictograph at Tadrart Acacus, southern west Libya. Journal of Cultural Heritage, 7, 344–
349.

Nelson, D. E. (1993). Second thoughts on a rock-art date. Antiquity, 67, 893–895.

Nelson, D. E., Chaloupka, G., Chippindale, C., Alderson, M. S., & Southon, J. R. (1995).
Radiocarbon dates for beeswax figures in the prehistoric rock art of northern Australia.
Archaeometry, 37, 151–156.

Pettitt, P., & Bahn, P. (2003). Current problems in dating Palaeolithic cave art: Candamo
and Chauvet. Antiquity, 77, 134–141.

Pettitt, P., & Bahn, P. (2015). An alternative chronology for the art of Chauvet cave.
Antiquity, 89, 542–553. doi:10.15184/aqy.(2015).21

Pettitt, P., & Pike, A. (2007). Dating European Palaeolithic cave art: Progress, prospects,
problems. Journal of Archaeological Method and Theory, 14(1), 27–47.

Plagnes, V., Causse, C., Fontugne, M., Valladas, H., Chazine, J.-M., & Fage, L.-H. (2003).
Cross dating (Th/U-14C) of calcite covering prehistoric paintings in Borneo. Quaternary
Research, 60, 172–179.

Quiles, A., Valladas, H., Bocherens, H., Delqué-Koliˇc, E., Kaltnecker, E., van der Plicht,
J., … Geneste, J-M. (2016). A high-precision chronological model for the decorated Upper
Paleolithic cave of Chauvet-Pont d’Arc, Ardèche, France. Proceedings of the National
Academy of Sciences 113(17), 4670–4675. doi:10.1073/pnas.1523158113

Ridges, M., Davidson, I., & Tucker, D. (2000). The organic environment of paintings on
rock. In G. K. Ward & C. Tuniz (Eds.), Advances in dating Australian rock-markings (pp.
61–70). Occasional AURA Publication 10. Melbourne: Australian Rock Art Research
Association.

Rowe, M. W. (2009). Radiocarbon dating of ancient rock paintings. Analytical Chemistry,


81, 1728–1735.

Page 19 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

Ruiz, J. F., Hernanz, A., Armitage, R. A., Rowe, M. W., Viñas, R., Gavira-Vallejo, J. M., &
Rubio, A. (2012). Calcium oxalate AMS 14C dating and chronology of post-Palaeolithic
rock paintings in the Iberian Peninsula. Two dates from Abrigo de los Oculados
(Henarejos, Cuenca, Spain). Journal of Archaeological Science, 39(8), 2655–2667.

Russ, J., Hyman, M., Shafer, H. J., & Rowe, M. W. (1990). Radiocarbon dating of
prehistoric rock paintings by selective oxidation of organic carbon. Nature, 348, 710–711.

Russ, J., Hyman, M., Shafer, H. J., & Rowe, M. W. (1991). 14C dating of ancient rock art: A
new application of plasma chemistry. Plasma Chemistry and Plasma Processing, 11(4),
515–527.

Sand, C., Valladas, H., Cachier, H., Tisnérat-Laborde, N., Arnold, M., Bolé, J., & Ouetcho,
A. (2006). Oceanic rock art: First direct dating of prehistoric stencils and paintings from
New Caledonia (Southern Melanesia). Antiquity, 80, 523–529. doi:10.1017/
S0003598X0009400X

Schiffer, M. B. (1987). Formation processes of the archaeological record (pp. xxii + 428).
Albuquerque: University of New Mexico Press.

Steelman, K. L., Carrera Ramírez, F., Fábregas Valcarce, R., Guilderson, T., & Rowe, M.
W. (2005). Direct radiocarbon dating of megalithic paints from north-west Iberia.
Antiquity, 79, 379–389.

Steelman, K. L., & Rowe, M. W. (2012). Radiocarbon dating of rock paintings:


Incorporating pictographs into the archaeological record. In J. McDonald & P. Veth
(Eds.), A companion to rock art (pp. 565–582). Chichester, UK: John Wiley & Sons. doi:
10.1002/978111825(3892).ch32

Taçon, P. S. C., Boivin, N., Petraglia, M., Blinkhorn, J., Chivas, A., Roberts, G., … Zhao, J. -
X. (2013). Mid-Holocene age obtained for nested diamond pattern petroglyph in the
Billasurgam Cave complex, Kurnool District, southern India. Journal of Archaeological
Science, 40, 1787–1796.

Taçon, P. S. C., May, S. K., Fallon, S. J., Travers, M., Wesley, D., & Lamilami, R. (2010). A
minimum age for early depictions of southeast Asian praus in the rock art of Arnhem
Land, Northern Territory. Australian Archaeology, 71, 1–10.

Taylor, R. E., & Bar-Yosef, O. (2014). Radiocarbon dating: An archaeological perspective


(2nd ed.). Walnut Creek, CA: Left Coast Press.

Valladas, H. (2003). Direct radiocarbon dating of prehistoric cave paintings by


accelerator mass spectrometry. Measurement Science and Technology, 14, 1487–1492.

Page 20 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

Valladas, H., & Clottes, J. (2003). Style, Chauvet and radiocarbon. Antiquity, 77, 142–145.

Van der Merwe, N. J., Sealy, J., & Yates, R. (1987). First accelerator carbon-14 date for
pigment from a rock painting. South African Journal of Science, 83(1), 56–57.

Vázquez, C., Maier, M. S., Parera, S. D., Yacobaccio, H., & Solá, P. (2008). Combining
TXRF, FT-IR and GC-MS information for identification of inorganic and organic
components in black pigments of rock art from Alero Hornillos 2 (Jujuy, Argentina).
Analytical and Bioanalytical Chemistry, 391, 1381–1387.

Watchman, A. (1999). A universal standard for reporting the ages of petroglyphs and rock
paintings. In M. Strecker & P. Bahn (Eds.), Dating and the earliest known rock art (pp. 1–
3). Oxford: Oxbow.

Watchman, A. L. (1987). Preliminary determinations of the age and composition on


mineral salts on rock art surfaces in the Kakadu National Park. In W. R. Ambrose & J. M.
U. Mummery (Eds.), Archaeometry: Further Australasian studies (pp. 36–42). Canberra,
AUS: Australian National University.

Watchman, A. L. (1993). Evidence of a 25,000 year old pictograph in northern Australia.


Geoarchaeology, 8, 465–473.

Watchman, A. L., & Campbell, J. (1996). Microstratigraphic analyses of laminated oxalate


crusts in northern Australia. In M. Realini & I. Toniolo (Eds.), Oxalate films in the
conservation of works of art. Proceedings of the 11th International Symposium. Milan
(pp. 409–422). Bologna: Editeam.

Watchman, A. L., & Cole, S. (1993). Accelerator radiocarbon dating of plant-fibre binders
in rock paintings from north-eastern Australia. Antiquity, 67, 355–358.

Watchman, A. L., David, B., McNiven, I. J., & Flood, J. M. (2000). Micro-archaeology of
engraved and painted rock surface crusts at Yiwarlarlay (the Lightning Brothers site),
Northern Territory, Australia. Journal of Archaeological Science, 27, 315–315.

Watchman, A. L., & Jones, R. (2002). An independent confirmation of the 4ka antiquity of
a beeswax figure in Western Arnhem Land, northern Australia. Archaeometry, 44(1), 145–
153.

Whitley, D. S. (2012). In suspect terrain: Dating rock engravings. In D. McDonald & P.


Veth (Eds.), A companion to rock art (pp. 605–624). Chichester, UK: John Wiley & Sons.
doi:10.1002/978111825(3892)

Page 21 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017


Radiocarbon Dating in Rock Art Research

Notes:

(1) This value is further qualified by an associated age offset because of the presence of a
possible dead carbon fraction (see text). Taçon et al. (2013:1792) suggest a conservative
terminus ante quem of circa 4100 cal BP.

Fiona Petchey
Radiocarbon Dating Laboratory, University of Waikato

Page 22 of 22

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University College London; date: 24 March 2017

You might also like