Sei sulla pagina 1di 17

INSTITUTE OF PHYSICS PUBLISHING

MEASUREMENT SCIENCE AND TECHNOLOGY

Meas. Sci. Technol. 14 (2003) 14931509

PII: S0957-0233(03)56707-8

Use of luminescence dating in archaeology


James K Feathers
Luminescence Dating Laboratory, University of Washington, Box 353100, Seattle,
WA 98195-3100, USA
E-mail: jimf@u.washington.edu

Received 27 November 2002, in final form and accepted for publication


23 April 2003
Published 29 July 2003
Online at stacks.iop.org/MST/14/1493
Abstract
With improvements in methodology and instrumentation, luminescence
dating is becoming a much more useful chronometric tool in archaeology.
Procedures for dating ceramics are relatively routine and their accuracy has
been demonstrated in a number of studies. Research is aimed at applying
ceramic dating in situations where other methods lack the direct dating
ability of luminescence to resolve chronological problems. Sediment dating
is more difficult because of uncertainty in the extent of zeroing or because of
mixtures of different aged deposits. These problems have been addressed by
isolating signals most likely to be zeroed and by dating single aliquots, and
ultimately single grains. Single-grain dating now has the potential not only
to identify and date mixed deposits but to provide valuable information on
site formation processes. This is particularly critical for judging site
integrity at controversial sites.
Keywords: luminescence dating, ceramics, sediments, archaeological

method, single grains


1. Introduction
Until very recently archaeologists, particularly American ones,
have looked upon luminescence dating with some scepticism
(Feathers 2000a), which is somewhat ironic because it was
developed in an archaeological context, in Europe in the
1960s and 1970s, as a method for dating heated materials,
primarily ancient ceramics. And although the dating context
largely shifted to geological questions in the 1980s and 1990s
when application was extended to buried sediments, dating
of sediments at archaeological sites nonetheless constituted
a significant portion of this work. Still lingering doubts
about accuracy and reliability have remained. In the last few
years, however, the perception has begun to change. More
archaeologists have been looking at it as a dating option,
rivalling radiocarbon in some cases. This surge in popularity
can be measured in the United States by increased funding from
grant sources such as the National Science Foundation, the
inclusion of luminescence dating in archaeological contracts
by government agencies and the proliferation of the number
of laboratories. (Ten at the last count in the United States,
compared with only five a few years ago.) Similar expansions
are occurring elsewhere.
This sudden interest stems in no small part from both
methodological advances, particularly in optically stimulated
0957-0233/03/091493+17$30.00 2003 IOP Publishing Ltd

luminescence (OSL), and improvements in instrumentation.


Both have made luminescence dating more accurate and more
precise. But it also must arise from an increased appreciation
for the intrinsic value for archaeology of luminescence dating,
the ability to provide direct calendar dates for archaeological
events of interest. This paper reviews recent advancements in
instrumentation, method and application that have increased
the value of luminescence dating for addressing archaeological
questions.

2. Background
Luminescence is the emission of light from crystalline
materials following the absorption of energy from an external
source. It is distinguished from other light emissions such
as fluorescence by a time interval between absorption and
emission, by a duration in a metastable energy state, that
permits use for dating. The external source of energy can
take a variety of forms, but for dating purposes the source
is naturally occurring, ionizing radioactivity (alpha, beta,
gamma and cosmic radiation). Release of the absorbed
energy requires a stimulus, which again can have varied
sources, but for materials used in dating the source is usually
either heat, resulting in thermally stimulated luminescence, or

Printed in the UK

1493

J K Feathers

Figure 1. A simplified schematic of the physical process behind luminescence dating. At point A all traps (e.g. minerals in the raw material
of the pottery) have been filled. Between points A and B, a zeroing event (heating of pottery) empties the traps and in the process
luminescence is emitted (unnoticed, of course, by the potter). Through exposure to natural radiation the traps are gradually filled through
time, a process shown here as linear although it may not be at all points in time, particularly as saturation is neared. At point C to point D,
the traps are again emptied, this time in the laboratory while luminescence is measured. The measured signal is proportional to the amount
of trapped charge. If one knows how much radiation is required to arrive at point C, a value called the equivalent dose, then dividing by the
average dose rate will yield the time from B to D.

more commonly thermoluminescence (TL), or light, resulting


in OSL. The latter includes stimulation not only from the
visible region of the electromagnetic spectrum but from
infrared and ultraviolet radiation as well. Luminescence is
mainly observed in ordered crystals, the metastable levels
associated with crystalline defects that create localized charge
deficiencies. Quartz and feldspar are two minerals with
suitable luminescence properties commonly used in dating.
Luminescence is explained by reference to solid state
energy band theory (Aitken 1985, 1998, McKeever and Chen
1997). Ionizing radiation excites electrons from the valence
band, or ground state, across an energy gap to the conduction
band. In ideal crystals energy levels within the gap cannot
be occupied, but defects allow metastable energy levels within
the forbidden zone which can trap excited electrons or electron
vacancies (holes). Traps are emptied of these charge carriers
by stimulus from additional energy such as heat or light.
The probability of escape, p, is governed by an Arrhenius
relationship, which for TL is expressed as
p = se(E/kT )
where E is the trap depth or the amount of energy required
to move an electron from the trap to the conduction band (or
a hole to the valence band), k is Boltzmanns constant and
T is temperature. For OSL additional terms related to the
intensity of light and the ability of the trap to absorb particular
wavelengths are required. Upon escape the charge carrier may
recombine with opposite charge residing at deeper traps called
recombination centres and in the process release energy in the
form of luminescence.
Materials with suitable luminescence properties can be
dated because at some point in the past traps are emptied
of their charge by sufficient exposure to heat or light. This
amounts to a zeroing event. If measurement of luminescence
was made immediately after this zeroing, no signal would
be observed. Subsequently, traps become refilled because of
continued ionization by radioactivity and a latent luminescence
1494

signal steadily accumulates. When luminescence is measured


in the laboratory, this natural signal will be proportional in
size to the time since the last zeroing event (figure 1). The age
equation is written as a simple ratio:
Age (ka) = D E (Gy)/D R (Gy/ka)
where D E is equivalent dose in grays (the unit for absorbed
dose) and D R is the average dose rate over time. Time is given
as ka (1000 years) but other units of time could obviously be
substituted.
Equivalent dose, a concept sometimes difficult for
non-specialists to grasp, is essential for archaeologists to
understand if they are to evaluate luminescence dates, since
it is often the term with the greatest uncertainty. It is defined
as the amount of radiation, in terms of absorbed dose, required
to produce a luminescence signal equivalent to the natural one
measured on the sample. In other words, how much radiation
is needed to get from zero luminescence to the current,
natural luminescence. Equivalent dose (D E ) is measured by
calibrating the natural signal against laboratory-administered
radiation, which acts as a proxy for natural radiation.
While the bulk of luminescence research is directed at
accurate estimates of D E , this is not to underestimate the equal
importance of an accurate estimate of the dose rate. The current
dose rate is often assumed to represent the average dose rate
because of the long half-lives, of the order of 109 years, of the
major sources, 40 K, 238 U and 232 Th. The dose rate includes
an internal component (short-ranged alpha and beta radiation)
from the sample itself and an external component (long-ranged
gamma and cosmic radiation) from the environment. The
external dose rate can be difficult to estimate because of
compositional inhomogeneity in the immediate environment
(so-called lumpy environmentsSchwarcz (1994)). Also
problematic is a potential change in dose rate, either internal
or external, through time either from a simple change in the
amount of overburden or from a change due to other geological
processes that add or subtract radionuclides. Such changes

Use of luminescence dating in archaeology

can often be detected and sometimes corrected by observing


stratigraphic trends (Mercier et al 1995a), by documenting the
presence of disequilibrium in the uranium and thorium decay
series (Readhead 1987, Olley et al 1996), or by comparing
equivalent dose among different minerals or different grain
size fractions which experience differential dose rates (Vogel
et al 1999, Feathers 2002).
TL is usually measured by heating a sample aliquot at a
steady rate and detecting the light emission with a sensitive
photomultiplier. As the temperature increases, the probability
of escape from traps increases while the number of electrons
remaining in the traps decreases. This creates a peak for any
particular kind of trap. The usual outcome is a composite
of peaks, relating to different traps, called a glow curve. An
alternative method of measurement is to raise the temperature
rapidly to some point and hold it there, producing a decay
curve (called isothermal TL), which again is often a composite
from several traps. OSL is usually measured as a decay curve,
sometimes called a shine down, when the sample is illuminated
with constant power.
Because traps have different energy depths, some traps are
not very stable at ambient temperatures and lose their electrons
over the time period relevant to archaeology. Other traps
are sufficiently deep that they are not easily emptied during
the depositional event. Obviously, neither very shallow nor
very deep traps will be useful for dating, and an important
task in dating is to isolate those traps which were emptied at
the time of interest and which have been stable since then.
Since the natural signal is a product of only traps that are
stable over the time of interest, and since signals induced
by artificial irradiation will include unstable components, the
task involves making these signals comparable, using various
thermal (preheat) treatments to simulate long periods of time
and plateau tests to gauge comparability.

3. Archaeological significance
Luminescence dates past exposure to heat and light, and
because these exposure events are often the same events
archaeologists are interested in, luminescence dating has a
strong advantage over other dating methods. Dean (1978)
argued the importance, when interpreting dating results, of
distinguishing dating events from target events. In tree-ring
dating, for example, the dating event is the time when the tree
was cut down, but the target event might be the use of the
tree limb in construction. The two events might be widely
separated in time, as, for example, in the cases of stockpiled
wood or use of naturally fallen timbers, the so-called old wood
effect (Schiffer 1986). Bridging arguments are then needed
to associate the dating event with the target event, arguments
often required for interpreting dates from the most commonly
used chronometric method, radiocarbon dating.
In luminescence dating, the target events and the dating
events are often the same events, obviating the need for
bridging arguments. In the case of ceramics, for example,
the dating event is the last exposure to sufficient heat to zero
the signal. Firing temperatures of at least 500 C are normally
required, although temperatures as low as 300 C are sufficient
if the duration of heating is long, and commonly the last such
exposure was when the ceramic was made or in some cases

when the ceramic was used, events that correspond precisely


with what archaeologists are interested in. The spectre of postdepositional heating, such as naturally occurring forest fires or
modern agricultural practices, is often raised but is rarely a
problem. Our laboratory has been involved in dating ceramics
from the North Carolina coastal plain, an area that has been
subject to frequent forest fires, but we nevertheless have been
able to achieve a reasonable chronology (Herbert et al 2002).
We are currently dating ceramics from central Amazonia, an
area subject to modern slash and burn agriculture, to see how
much of a problem that might create.
Heated lithics are somewhat different. While heat
treatment was commonly employed in prehistory to improve
flaking qualities, such heating was often not sufficient to reset
the signal. The manufacturing process may be directly dated
if overheated lithics (wasters) can be located (Wilhelmsen and
Miles 2003), but more often luminescence dating addresses a
discard event, when lithic tools became accidentally burned in
hearths. Use of the hearths then becomes the dating event, an
event also of interest to archaeologists, and one that also can
be directly addressed by dating hearth stones, again providing
the fire was hot enough and the rocks thin enough to counter
insulating effects (Anthony et al 2001).
For sediments the dating event is the last exposure to
sufficient sunlight. This is a depositional event. Unlike the
case of ceramics, however, where sufficient heat is nearly
always applied in manufacture (there are rare exceptions for
very low-fired wares), the depositional agent of sediments may
not allow sufficient exposure to zero the signal fully. The
signal may not be zeroed at all, so that a derived age would
actually refer to some previous depositional event, or it may
be partially zeroed, retaining a sufficient residual from time
elapsed since the previous, full-zeroing event. For example,
some colluvial depositional agents such as landslides may not
expose sediment grains at all to light. High-energy fluvial
regimes, such as glacial outwash, because of rapid turbidity
and high sediment loads may result in only partial exposure
or differential exposure of grains (Gemmell 1997). Aeolian
derived sediments are usually considered the best candidates
for luminescence dating because airborne time allows ample
sun exposure. Sufficient exposure, however, is governed by
local circumstances not by any hard and fast rules about
depositional regimes. Our laboratory has had experience of
fluvial sediments having been well exposed and nearby aeolian
ones poorly exposed (Feathers 2003).
The fact that sediments are composites of uncemented
grains also makes them more problematic for dating than
discrete objects like ceramics or lithics. A single deposit can
result from different transport processes, some of which may
permit wider exposure than others. Post-depositional mixing
from turbation or deflation is also a major problem in dating
sediments (Bateman et al 2003).
Despite these problems, which are becoming more
resolvable by methodological developments, sediment dating
has high potential for archaeological chronology. It provides a
direct date for the deposition of artefact-bearing stratigraphic
layers, reducing reliance on the assumption that materials in
the deposits (such as charcoal) are the same age as the deposit.
Luminescence may also allow dating of anthropogenic
sediments. Artificial construction (walls, canals, mounds and
1495

J K Feathers

other earthworks) often has involved sedimentary material,


humans being the transport agent (Stein 1987). If the material
was exposed during construction, luminescence may provide a
means of directly dating such episodes (e.g. Bush and Feathers
2003, Feathers 1997), or if the construction resulted in pits or
trenches, luminescence may address abandonment by dating
the immediate fill (Lang et al 1999, Sanderson et al 2003).
The discussions that follow are not a review of
luminescence work in archaeology but a selection of examples
that illustrate particular applications. I have tried to select
examples that postdate the exhaustive review of Roberts
(1997). The examples are weighted to some degree towards
our work at the University of Washington laboratory, only
because of the authors familiarity. This literature can be well
sampled by reviewing issues of Radiation Measurements and
Quaternary Science Reviews, two journals that publish many
luminescence applications, and by checking bibliographies
published regularly in Ancient TL.

4. Dating ceramics
Luminescence dating first was developed in the context of
dating ceramics, using TL. The basic procedure was worked
out by the late 1970s mainly by the Oxford laboratory under
the leadership of Martin Aitken (Aitken 1985, Feathers 2000a).
Even after the development of OSL for dating sediments
(Huntley et al 1985), TL has continued to be employed for
dating ceramics. The main advantage of OSL in sediment
dating, the ease in isolating the most light-sensitive traps, is not
a problem for ceramics. Still, OSL may have some advantages
for ceramics in terms of requiring a smaller sample size, but
limited work has been done (Barnett 2000a, Guibert et al 2001,
and Yurdatapan 2000), and often ceramic
Hong et al 2001, Oke
materials produce weak OSL signals compared with TL signals
(Hong et al 2001) and may behave differently from unheated
material (Barnett 2000a).
Luminescence dating commonly employs either polymineral fine grains or coarse silt- to sand-sized grains of either
quartz or feldspars. Using coarse grains has the advantages
of eliminating the influence of short-ranged alpha irradiation, since neither quartz nor feldspar have appreciable internal sources of alpha irradiation, and of involving only a single mineral with fairly well-understood luminescence properties. Coarse-grain dating is still commonly used for ceramics
(e.g. Barnett 2000a), but many ceramics lack abundant inclusions to make this practical. Our laboratory usually relies on
fine grains. While having to include alpha irradiation in the
dose rate and the complications that arise because of the lower
efficiency at trapping charge by alpha as opposed to beta and
gamma radiation, fine grains can turn this to an advantage in
that less reliance is put on the more uncertain external dose
rate.
Two general techniques have been employed to measure
D E by TL (figure 2). Both involve using several aliquots of
sample material, some of which are used to measure the natural
signal and others of which are given calibrating irradiations.
With the first, called additive dose, incremental irradiations
are given to different aliquots that still retain their natural
dose and a growth curve is constructed plotting irradiation
against luminescence signal. The natural signal forms the
1496

lowest point on this curve, which is then extrapolated back


to zero dose to estimate D E . Such a procedure fails to take
into account changes in the shape of the growth curve in the
extrapolated region, and consequently additive dose, especially
for younger samples, frequently underestimates the age. The
second technique, called regeneration, applies incremental
irradiations to aliquots that have first been measured for their
natural signal and thus zeroed. This procedure regenerates
the growth curve from zero and the natural signal is fitted
into the curve by interpolation, the shape of the growth
curve not being as much of a problem. However, heating
of the sample to measure the natural signal often changes the
luminescence sensitivity of the sample (i.e. it now takes more or
less irradiation to produce the same amount of luminescence),
so that the regeneration curve no longer mimics the original
growth curve of which the natural signal is a part. Interpolation
then leads to gross errors.
The traditional solution to the problems of both techniques
has been to use the regeneration growth curve intercept to
correct the extrapolated value from additive dose (the so-called
supralinearity correction). Our laboratory has adopted the
more precise slide technique, the idea for which was first
advanced by Valladas and Gillot (1978) and Readhead (1988),
and then developed for sediment dating by Prescott et al (1993)
and Huntley et al (1993). Both additive dose and regeneration
curves are constructed. If the curves are the same shape (which
they almost always are), then the regeneration curve can be
corrected for sensitivity change by applying a multiplicative
factor to make the two curves parallel, a procedure that may
not be valid for samples close to saturation (almost never the
case for ceramics). Once this correction is made, the horizontal
distance (along the dose axis) between the curves can be taken
as the D E . Our laboratory has dated over 500 ceramics in this
manner and has produced results that agree with independent
evidence (where such is available) more than 85% of the time,
perhaps higher if the independent evidence, whose accuracy is
often unknown, were truly accurate (e.g. Feathers 2000b).
Tests for accuracy involve evaluation of dates for events
that have been securely dated by independent means. This
is not straightforward because seldom are the dating events
precisely the same for luminescence and the independent
method, even if the target events are the same. An example
worth detailing from our laboratory has been a comparison
of TL ceramic dates with tree-ring dates from several shortoccupation Navajo sites in northwest New Mexico (Dykeman
et al 2002, Feathers 2000b). From 12 sites both TL and treering dates were available. Of the 15 TL dates, only 27% were
within one standard deviation of the tree-ring dates, but a more
telling comparison was the aggregate data. Figure 3(a) shows
histograms of the two sets of dates. Each TL date is treated
as a normal probability distribution and the intervals in the
histogram reflect summing of probabilities of any one date
falling in that interval. (The TL date distribution will have
a broader distribution because of poorer precision than the
1 year precision of tree-ring dating.) The distributions are
similar with a main mode around AD 1700. An earlier minor
mode apparent in the tree-ring data is less defined in the TL
distribution. However, if the TL distribution is divided into the
two ceramic types represented (figure 3(b)), the bimodality of
one of the types closely matches the bimodality in the treering dates, while both types contribute to the main mode. The

Use of luminescence dating in archaeology

(a)

(b)

(c)

Figure 2. (a) Additive dose growth curve for sample UW736 (pottery from southern Arizona). Dose refers to artificial irradiation given in
the laboratory, so that the natural signal is located at the zero point. A linear fit to the points yields an extrapolated dose value of about
0.8 Gy. (b) Regeneration growth curve for the same sample. The natural signal is represented by the triangle and interpolating that value
into the linear fit of the points yields a dose value of about 1.15 Gy. (c) Both growth curves after they have been shifted horizontally into
coincidence with the best fit shown. The ratio of the two slopes is about 1.2 and this is used to correct for change in sensitivity. The amount
of the shift (or slide) for the corrected points is about 1.0 0.08 Gy, taken as the best estimate for D E .

75th probability distribution of the TL dates for both ceramic


types (AD 16401780 for Gobernador Polychrome and AD
15201740 for Dinetah Gray) is remarkably close to the
range of production for both types estimated from radiocarbon
and tree-ring dates from throughout the region (AD 1630
1775 for Gobernador Polychrome and AD 15401800 for
Dinetah Gray) (Dykeman et al 2002). This underscores
the problem of comparing two techniques that date different
events. The tree-ring dates address cut dates (at best) for
the trees ultimately used in construction, while the TL dates
address ceramic production. The direct comparison show
less correspondence because the events dated occurred at
different times, the distance between them being in part a
function of occupation duration. The aggregate comparisons
reflect regional activities over a block of time, where temporal
differences in construction/production events may average out.
Here the correspondence is much closer. This work also
demonstrates that several dates (TL or otherwise) are necessary
to flesh out a chronology. One or two dates (of any kind)
are not useful beyond a rough chronological indicator. The
accuracy of this work is not an aberration but merely adds to
the growing evidence of the accuracy of TL dating applied
to ceramics (e.g. Barnett 2000a, Godfrey-Smith et al 1997,

Guibert et al 1994, Kojo 1991, Sampson et al 1972, Whittle


and Arnaud 1975).
Precision is a separate problem. The lower resolving
power of TL compared with radiocarbon or tree-ring dating
is a function of the greater number of variables and intrinsic
uncertainties involved (although a more precise technique
is not necessarily a more accurate one, and TL retains its
advantage as a direct dating method). Our laboratory routinely
produces dates with better than 10% precision (Feathers
2000b). These error terms reflect mainly analytical error,
but also include systematic uncertainties on unmeasurable
quantities such as past moisture contents (which affect the
radiation dose the sample receives).
For many ceramic materials the fine-grain signal is
dominated by contributions from feldspar, which have much
higher luminescence sensitivity than other minerals such as
quartz. A persistent problem in dating feldspars is anomalous
fading, the non-thermal loss of stored charge from traps
that from kinetic considerations should be stable at ambient
temperature (Spooner 1994a). Anomalous fading has been
attributed to quantum mechanical tunnelling (e.g. Visocekas
et al 1994). By this theory a trapped electron has a small
but finite probability of being found outside the energy barrier
1497

J K Feathers

(a)

(b)

Figure 3. (a) Comparison of tree-ring dates with luminescence


dates. Each TL date is treated as a normal probability distribution
and the intervals in the histogram reflect summing of probabilities of
any one date falling in that interval. (b) The same TL data, except
divided between two ceramic types.

that retains it. If this probability distribution overlaps that


of a recombination centre, the electron can escape, even at
very low temperatures. Such fading appears ubiquitous among
feldspars, although some correlation has been found between
the amount of fading and the degree of disorder in the crystal
structure (Visocekas et al 1994). Feldspars of volcanic origin,
where an ordered structure is prevented by rapid cooling,
tend to have more problems with fading, which is one reason
ceramics from areas with predominant volcanic geology (such
as Pacific islands) are difficult to date. Such regions also tend
to have low radioactivity (Prescott et al 1982).
Fading can be detected by loss of signal as a function
of storage time and is routinely checked. Our experience is
that about 1 in 7 ceramics show evidence of serious fading.
Quantum tunnelling predicts that the rate of fading will be
proportional to 1/t where t is some unit of time, and therefore
the number of remaining trapped electrons decreases linearly
with ln(t). This is because centres and traps that are close
together will fade first and as these are depleted the rate slows.
This logarithmic relationship has allowed development of a
correction procedure that appears to work for samples that
are neither very young nor old enough that most fading has
1498

been completed (Huntley and Lamothe 2001). The procedure


uses fading rates measurable in the laboratory to estimate
a percentage decrease through an interval of log time and
through iteration calculates an age that with the estimated
fading rate would result in the derived, faded age. This
procedure has been tested on coarse feldspar grains using
infrared stimulated luminescence (IRSL) in sediment dating
(Huntley and Lamothe 2001), but our laboratory has tried using
it for fine-grained ceramics using TL. Although a systematic
study remains to be done, initial results seem to indicate that
it works, although at a cost in precision (figure 4).
Other procedures have been proposed to deal with
anomalous fading, mostly in the context of sediment dating
but probably applicable to heated materials as well. Lamothe
and Auclair (2000) have found variable fading rates among
feldspar grains from the same deposit and have used the
relationship between equivalent dose and fading rate among
grains to extrapolate to an equivalent dose with zero fading.
Recent research in both TL and IRSL have identified a far-red
to near-infrared emission band from feldspars that does not
seem to exhibit fading (Zink and Visocekas 1997, Fattahi and
Stokes 2003). Isolating a red signal requires special attention
to photomultiplier and filter set-up to minimize red oven glow,
in the case of TL, and interference from the stimulating light,
in the case of IRSL. For ceramics, perhaps the simplest way to
circumvent fading is to extract quartz grains for dating. Since
for many ceramics the amount of suitable quartz might be
small or difficult to isolate, further research into using OSL
for ceramics seems warranted.
Methodological research into ceramic dating has recently
become almost completely overshadowed by research into
sediment dating (but see Barnett 2000b, Guibert et al 2001).
Ceramic dating is generally considered routine, and most
laboratories are oriented towards dating sediments. The most
promising area of ceramic dating research in fact might be
application, where advantage of the direct dating capabilities
of luminescence dating can be taken. Unfortunately, little work
demonstrating this potential has been published, and the review
by Roberts (1997) was able to list applications that mainly
made substantive rather than methodological contributions to
archaeology.
4.1. Mixed assemblages
A common problem in archaeology is isolating different aged
components from artefact concentrations that have been mixed
by post-depositional processes (either natural or cultural) or by
virtue of being on an old surface. For example, pits of various
shapes and sizes are common features in many archaeological
sites. Even where pits show no evidence of disturbance or slow
infilling, misleading conclusions can result from assuming that
everything in the pit is the same age. Yet radiocarbon dating of
charcoal or other plant remains in pits is a common approach
to attempt dating ceramics found in the same pits.
Barnett (2000b) has argued that the larger the pit the more
likely it contains materials of different agesin terms of their
origin or manufactureeven if the infilling, or deposition, was
a single event. In smaller pits, the age of deposition is more
likely to correspond to the age of the contents. Using a series
of TL dates on Iron Age/Roman sites in Great Britain she

Use of luminescence dating in archaeology

Figure 4. Examples of two samples showing evidence of anomalous fading. The table beneath the figure compares the corrected age, using
the Huntley and Lamothe (2001) method, with expected ages derived from independent data.

showed that large enclosure pits contained ceramics dating


to a wide range. Dating of the infilling could be estimated,
short of dating the sediments themselves, by dating a large
number of ceramics and using Bayesian statistics to estimate
the most recent date. In contrast, smaller pits from other sites
contained ceramics whose dates were much more constrained,
especially after taking into consideration complications in the
environmental dose rate due to differential radioactivity inside
and outside the pit.
Holt and Feathers (2003) have shown that even small
pits can have problematic contents. The occurrence of two
ceramic types, previously thought to be separated in time, in
the same pits at the BaehrGust site in the Lower Illinois River
Valley has puzzled local archaeologists, leading some to argue
that the ceramics were in fact contemporary. This implied
that a regional trade and ceremonial network called Hopewell
extended much later in the Lower Illinois than elsewhere.
TL dating of pairs of ceramics from the same pit, however,
clearly showed that one ceramic type was hundreds of years
older than the other and that the radiocarbon dates from some
pits were highly misleading because they represented averages
of different aged charcoal. In this case, information on the
environmental dose rate was not even available, but since the
issue was the relative age between the two sets of ceramics,
both of which experienced the same environmental dose rate,
interval- rather than ratio-scale ages were obtained. By holding
several sources of possible systematic error constant (including
not only the environmental dose rate but also moisture contents
and calibration of the laboratory irradiation source), interval
dating improved precision by as much as 25%.
Pits are not the only problem. Many sites in North
America, for example, occur on erosional surfaces, so that
the entire record rests on the modern surface. Identifying
reoccupations is very difficult and radiocarbon is completely
useless because the dated organics cannot be comfortably
associated with any of the artefacts. Our laboratory is
addressing this issue at several sites in the Lower Mississippi
Valley by dating a large number of ceramics from them. This

work is still in progress. An issue is the suitability of surface


finds for dating because of uncertainties in the external dose
rate. Several studies have demonstrated the accuracy of dating
surface ceramics by TL (Dunnell and Feathers 1994, Sampson
et al 1997), suggesting the uncertainties from the external dose
rate are no greater than those from buried sites and perhaps less
so because of less uncertainty in the thickness and duration of
overburden.
4.2. Site duration
Duration of occupation in any one locus is an important
variable for understanding human demographic trends
and their impact on environment, economy and political
complexity (Varien and Mills 1997). Controlling assemblage
duration is also necessary when comparing similarities among
assemblages since variability will increase with duration.
Estimating duration, however, has been a difficult problem,
usually accomplished by an accumulation of radiocarbon or
tree-ring dates. The problems of association are as acute
here as in any other dating endeavour (e.g. Nash 1997), and
other archaeologists have attempted more direct measures such
as artefact accumulations, most parsimoniously expressed by
the Schiffer (1975) discard equation positing a relationship
between discard and occupation span (Varien and Mills 1997,
Gallivan 2002). While gross comparisons are possible, lack
of control over a number of variables often creates high
uncertainties that cannot be quantified. Direct dating of
artefacts by luminescence provides a viable alternative. Dating
several ceramic shards from one site will provide a direct
measure of duration of use that can be statistically quantified
(e.g. by using the OxCal program of Bronk Ramsey (1995)).
The more dates obtained, the higher the precision of site
duration. Precision can also be increased because only
interval-scale dating is required, as suggested earlier. Again,
applications of this sort have not been published, but our
laboratory is engaged in trying to determine the duration of
large late prehistoric towns in the Mississippi Valley.
1499

J K Feathers

4.3. Small sites and other difficult-to-date assemblages


Since the 1960s, at least, archaeologists have been concerned
with securing a representative sample of the archaeological
record. This has meant attention paid not just to large,
conspicuous concentrations of artefacts but also to smaller
concentrations that dot the landscape. Where traditional
studies relied on large-scale excavations, the representative
approach has necessarily relied on surface survey. In most
places such survey has yielded numerous small artefact
concentrations. While important for understanding land use
patterns, these small sites have not attracted attention because
they often lie completely within the ploughzone and are
difficult to date by traditional means. Small artefact samples
reduce the probability of recovering time diagnostic artefacts
and organics, if preserved, cannot be associated with other
artefacts (Beck 1994). Without knowing how these sites fit
into a local cultural chronology, it is difficult to understand
their significance. Our laboratory is beginning to address this
problem by TL dating ceramics from small sites in several
areas in the Mississippi Valley, where late prehistoric large
towns are surrounded by numerous small sites, sometimes
called farmsteads. A presumption has been that the large
sites represent nucleation of population from the earlier small
sites, but no one actually knows how old the small sites really
are. Luminescence dates are providing the first indication.
Luminescence dating of ceramics has also proved useful in
areas where other dating techniques are simply not applicable.
Our laboratory has been dating ceramics from the sandhills
of coastal North Carolina, an area where almost no charcoal
associated with archaeological remains has been preserved and
the chronology has been poorly developed. Using TL, we
have worked out a tentative chronology that reasonably follows
technological trends and that is consistent with chronological
work in nearby areas (Herbert et al 2002). Barnett (2000a) has
published TL dates of ceramics from Late Bronze Age and Iron
Age sites in England, a period of time when lack of stratigraphy,
few diagnostic pottery styles and a relatively flat 14 C calibration
curve have hampered development of a chronology. Her dates
agree reasonably well with independent evidence. She also
notes that chronologies based on technological properties of
the pottery, what she calls fabric and includes non-plastic
inclusions (temper), do not have much resolution. While
archaeologists have considered changes in temper during this
period in sequential terms, the TL dates suggest considerable
temporal overlap. Similarly, our laboratory has found that
the supposed temporal sequence of limestone, grog and shell
temper in the southeastern United States is not an accurate
portrayal. Limestone and grog tempers persist for a long period
after shell temper is introduced (unpublished data).

5. Lithics
The direct dating advantage of luminescence dating of
ceramics applies to lithics as well and methodological
considerations are similar. The substantive contributions of
lithic dating, however, are better known, so only briefly
mentioned here, largely because the use of lithics extends
well beyond the chronological range of radiocarbon dating.
In Palaeolithic studies, TL dating of lithics is often only
1500

one of a few dating options available. TL has particularly


been important in documenting the middle Palaeolithic and
transitions before and after (e.g. Mercier and Valladas 1994,
Mercier et al 1995b, 1999, 2000, Valladas 1992, Valladas et al
1998, 2001, Richter et al 2000, Porat et al 2002). Some lithics
have little internal sources of radiation, so dating requires
reliance on good estimates of the external dose rate, which may
include a high contribution from cosmic radiation (Richter et al
2000). This will become particularly critical if lithic dating is
extended to surface finds.
Of potentially wide applicability is using OSL or IRSL
to date the light-exposed surfaces of rocks, lithic artefacts
or stone architecture from archaeological sites. The dating
event is the last exposure to light of these materials, and
could refer to a burial event or, perhaps more interesting, to
the underside of lithic materials found on the surface. This
work is only beginning and will require special attention to
both collection procedures that prevent exposure (including
excavating at night, Morganstein et al (2003)) and development
of new measurement techniques and instrumentation to sample
surfaces (Habermann et al 2000, Greilich et al 2002, Theocaris
et al 1997).
This discussion on ceramics and lithics cannot be complete
without mentioning that luminescence may have additional
value for archaeology beyond dating.
Luminescence
properties have been used as sourcing criteria for determining
where material for artefacts originated (Akridge and Benoit
2001, Hashimoto et al 2001) and to assess past thermal
treatments (Hashimoto et al 2001).
Mention also needs to be made of the application of
luminescence to other kinds of archaeological materials, such
as glass and metallurgical slag. Luminescence applications to
these materials have a long history (Aitken 1985), but recent
studies (e.g. Chiavari et al 2001, Gautier 2001, Haustein et al
2001), while promising, suggest much work is still needed
before such dating can be considered routine.

6. Sediments
Sediments are inherently more difficult to date than ceramics
because of the uncertainty of sufficient bleaching at deposition
and because of the uncertainty in the integrity of the deposits.
Heating unambiguously zeroes most ceramics, and the greater
chemical and erosional resistance of ceramics makes them
less susceptible to post-depositional change. With sediments,
only partial bleaching or no bleaching at all may occur at
the depositional time of interest and mixing of different-aged
sediments may occur either during or after deposition. In either
case, following a procedure similar to that for ceramics may
result in gross inaccuracies.
Two general methods have been employed to deal
with partial bleaching or mixing. The first is to separate
out components of the luminescence signal that are most
susceptible to bleaching from sunlight and base dating only
on them. This minimizes the problem of partial bleaching.
The second is to date individually aliquots containing a small
amount of material, ultimately single grains. The distribution
of ages among the aliquots or grains is used to identify mixing.
A number of studies (e.g. Bailey 2001, Huntley et al
1996, Duller 1997, Krbetschek et al 1997) have shown that

Use of luminescence dating in archaeology

Figure 5. (a) TL regenerative glow curve of coarse-grained quartz from a sediment sample from a Brazilian rock shelter. The full curve is
shown as a dashed curve, with the low-temperature peak reduced in scale. The slowly bleaching peak was obtained after an extended bleach
with green light. The rapidly bleaching peak is obtained by subtracting the slowly bleaching peak from the full curve (Feathers 1997).
(b) OSL decay from another quartz sample. The inset shows three exponential components that when summed form the solid line through
the data points in the main graph. The figure is modified from one presented by Bailey et al (1997).

the total luminescence signal from either quartz or feldspar is


a composite of several signals arising from different traps or
recombination centres. Traps can be isolated by manipulation
of the excitation source (heat or light), because of the different
energy required to release the stored charge from different
traps, while recombination centres can be isolated by selection
of emission wavelengths by optical filters.
OSL has gained popularity for dating sediments because
it samples traps that are more susceptible to bleaching than
TL. In quartz, for example, the high-temperature TL signal is
made up of a rapidly bleaching peak and a slowly bleaching
peak (Franklin et al 1992, 1995) (figure 5(a)). The latter not
only is the result of slowly emptying traps but also contains
an unbleachable residual. OSL is associated with only the
rapidly bleaching component (Spooner 1994b). While the
two components can be isolated in TL (Franklin and Hornyak
1990), it is much less time consuming to use OSL. Even
within OSL, however, there are components that bleach at
different rates (Bailey et al 1997) (figure 5(b)). Typically, OSL
is stimulated using a narrow bandwidth of light of constant

intensity. Depletion of the trapped charge produces what is


called a shine-down curve. If only one trap were involved,
the curve would be exponential in shape, but fitting studies
have shown that the measured curves are a combination of
at least three, and maybe as many as five, exponentials, each
associated with traps of different bleaching rates. To sample
only the most rapidly bleaching component, which has the
highest likelihood of being emptied during deposition, dating
studies normally only utilize the initial small fraction of the
curve, and use the latter part of the curve as a subtracted
background (Aitken and Xie 1992, Banerjee et al 2000).
Comparing the D E derived from different parts of the shine
curve is one procedure used to detect partial bleaching (Bailey
et al 2003). If the D E from latter parts of the curve, where
traps more difficult to bleach are being sampled, match those
from earlier parts of the curve, an argument can be made
for adequate bleaching at deposition. More recently, some
laboratories have experimented with altering the intensity
of stimulating light during the luminescence readout (Bulur
et al 2000). If the intensity is progressively increased from
1501

J K Feathers

Table 1. SAR protocol (Murray and Wintle 2000), taken from table 1 in Murray and Clemmensen (2001).
1.
2.
3.
4.
5.
6.
7.
8.

Dose (= 0 Gy if natural signal)


Preheat (e.g. 260 C for 10 s)
OSL (e.g. 60 s at 125 C, stimulation using blue light, emission in UV), gives L i
Fixed test dose (e.g. 5 Gy)
Cut heat (e.g. 160 C)
OSL (same as step 3), gives Ti
Repeat steps 16 for a range of regeneration doses bracketing the natural dose, a zero dose and a repeat (or recycle) point
Calculate L i /Ti for natural and regeneration points and construct regeneration curve to determine D E

zero to some maximum, called linear-modulated OSL, the


different components can be separated as peaks, making fitting
procedures mathematically more tractable (Agersnap Larsen
et al 2000, Singarayer and Bailey 2003). For feldspars, it
has been found that the most bleachable traps are sensitive to
infrared radiation. Stimulation by such a long wavelength is
explained by a resonance effect (800900 nm) coupled with a
thermal assistance mechanism (Hutt et al 1988, Aitken 1998).
IRSL has been widely used for dating feldspars, not only
because of rapid bleaching but because less attention is needed
to separate feldspars physically because few other minerals
have this effect. (Despite problems with fading, feldspars have
some important advantages over quartzhigher sensitivity and
higher saturation levelsmaking them particularly suitable for
older samples.)
Perhaps more commonly encountered are deposits that
have been mixed with differently aged grains either during
or after deposition. To separate these different populations
of grains, only aliquots with a small amount of material
on them are dated. Numerous studies have shown that the
percentage of grains of quartz that are actually sensitive to
luminescence (bright grains) is quite variable, between 5 and
10%, although sometimes as high as 40% (Duller et al 2000,
McFee and Tite 1998). Aliquots with a small amount of
grains, say 1020 grains, are likely to have only one or two
bright grains. If the deposit is made up of differently aged
populations, this should be detectable from the D E distribution
from several of these small aliquots, as evidenced by very
broad distributions or modality. A correlation of equivalent
dose with brightness is also characteristic of differently aged
or partially bleached samples, since brightness is a function
not only of luminescence sensitivity but also of age (Clarke
et al 1999, Li 1994). The best resolution obtains when dating
single grains, although requiring many more aliquots because
so many grains have no signal. To make such measurements
practical advances in both method and instrumentation were
required.
The method for determining D E outlined earlier for
ceramics requires multi-aliquots and is unsuitable for mixedaged samples because of the averaging effects of using several
aliquots. A single aliquot additive dose process was proposed
by Duller (1995). It involved measuring the natural signal and
successive signals from incremental doses, each preceded by a
preheat, by using OSL shines that were short enough that little
of the total signal was removed at each step, thus allowing
an additive dose build-up on the same aliquot. A correction
for loss of signal at each step, especially from the successive
preheats, was necessary. The method suffered not only from
uncertainties in the correction procedure, especially for older
samples that had begun to saturate and whose growth curves
1502

were not linear, but also from the inherent problem of additive
dosethe need to extrapolate to obtain D E and the subsequent
high dependence on achieving the correct shape to the growth
curve.
Dullers early work was not successful in devising a
single aliquot regeneration method, because of the problem
of sensitivity change. Once the sample had gone through one
cycle of irradiation, preheat and measurement, its sensitivity
changed, because of either changed luminescence efficiency
per unit trapped charge or changed rate of trap filling per
unit ionization, so subsequent signals from artificial doses
were not comparable. Murray and colleagues investigated
this problem for quartz and through a series of experiments
(Murray and Roberts 1998, Murray and Mejdahl 1999, Murray
and Wintle 2000) found that sensitivity changes could be
monitored by applying a small test dose after an OSL
measurement. The signal from the test dose was found to
have the same sensitivity as the preceding OSL signal and so
could be used as a correction to equalize the sensitivity of the
natural and subsequent regeneration signals. They devised
a procedure called single-aliquot regenerative-dose (SAR),
which constructs a growth curve from regeneration signals
whose sensitivity is corrected to match that of the natural
signal. The procedure was found to be relatively insensitive
to preheat temperature and size of the test dose. Because of
increased accuracy and precision, SAR is now the preferred
technique for most OSL applications involving quartz and is
being extended to feldspar and polymineral fine grains as well
(Wallinga et al 2000, Banerjee et al 2001, Auclair et al 2003,
Fattahi and Stokes 2003).
The general sequence of steps in SAR for a single aliquot is
given in table 1. The OSL measurement following the natural
or the regeneration dose is divided by the OSL measurement
following the test dose to equalize sensitivity. After several
regeneration cycles, the first regeneration dose is repeated to
make sure the correction procedure produces the same result
as the first time. A zero dose is usually included to assess
whether preheats or test doses are creating unwanted signals,
called recuperation. It is also recommended to recover a known
dose by first zeroing an aliquot, giving it a dose, treating that
dose as if it were the natural signal and then using the procedure
to see if the known dose can be recovered as the D E (Murray
and Wintle 2003). This ensures the procedure works for the
particular sample.
A variation of SAR was first applied to single grains of
quartz by Roberts et al (1998a, 1999) investigating a suspected
age overestimation from multi-aliquot analysis at Jinmium
rock shelter in Australia, which had placed human presence
in Australia by 116 ka, far earlier than previously believed.
By dating a large number of single grains, they were able

Use of luminescence dating in archaeology

to detect a complex distribution of ages, one part of which


suggested an age of less than 10 ka (agreeing with a radiocarbon
chronology), but also including a wide range of much older
ages. The older ages were attributed to grains that had eroded
off the rock shelter wall and never been properly zeroed by
sunlight. The younger ages were attributed to aeolian-derived
sediments that had been well zeroed and were believed to
represent the true depositional age.
The Jinmium work was done on conventional equipment,
and although labour intensive, the study benefited from a large
proportion of bright grains. For many samples, the proportion
of bright grains is much smaller and would have made the
study impractical. Using aliquots with just a small number of
grains, while increasing the likelihood of containing at least
one bright grain, loses some of the resolution provided by
single grains. In 1999 Ris Laboratories introduced a special
attachment to their basic luminescence reader that provided a
high-powered, finely focused laser that could stimulate single
grains (Btter-Jensen et al 2002). The grains are arranged
on discs containing a grid of 100 holes, one for each grain.
The discs can be irradiated and heated as a whole, while the
laser can rapidly measure the OSL from each grain. Daybreak
Nuclear, the other major supplier of luminescence equipment,
is devising a similar version. This allows the SAR sequence to
be performed on a large number of grains in a relatively short
time, depending on the irradiation lengths required. On some
Holocene-aged sediments recently dated in our laboratory, an
SAR sequence on 1000 grains could be completed in about
24 h (Feathers 2003).
Dating single grains is not as straightforward as the above
discussion suggests. Two problems make the interpretation
of D E distributions obtained from large numbers of grains
difficult. One is differential precision. Grains with the
brightest signals will produce values of D E with the greatest
precision (McCoy et al 2000). Even after eliminating grains
with no or very low signals barely above background, a range
of ages will be obtained even from a single-aged population
because of this differential precision. Statistical display
methods, such as histograms, which do not take this into
account, may be highly misleading. Galbraith et al (1999)
introduced the use of radial graphs, which plot distributions as
a function of precision, as a remedy (figure 6). The other
problem is microdosimetry. The three major components
of terrestrial radiation (gamma, beta and alpha rays) are
distinguished in part by their effective ranges. Alphas are short
ranged but of little consequence for coarse-grained quartz or
feldspar. Gammas are long ranged and affect all grains within a
sample about equally. Betas, however, have a maximum range
of about 2 mm. At the scale of single grains, the sources of
beta irradiation are not likely to be evenly distributed, so that
some grains will receive a higher beta dose rate than others.
Measuring the differential rates at such a small scale is not
currently possible, so the dose rate forming the denominator of
the dating equation is only an average for all grains. Averaging
will underestimate the age for some grains and overestimate it
for others.
Errors arising from differential precision and microdosimetry should be random, so that a single-aged population
of single grains should have a distribution in D E that is normal, with the mean representing the best estimate. This has

Figure 6. Radial graph for an aeolian sample from Mustang Springs


in west Texas. Precision on the x-axis is the relative standard error.
The y-axis is the number of standard deviations of individual
equivalent dose values from single grains away from some reference
value. The solid circles are those values within two standard
deviations of the reference. This is referred to here as the mean, but
is not the mean of the distribution but rather that value that
maximizes the number of points within two standard deviations. A
characteristic of a radial graph is that a radial line drawn from the
origin will pass through points of equal equivalent dose but with
different precision. The radial line bisects the scale on the right at
that equivalent dose value.

been the working assumption, and the challenge is to derive


ages from other distributions: multimodal, skewed or otherwise non-normal. For fluvial sediments, Olley et al (1999)
hypothesized that because moving water entraps grains from
different sources of different ages and can attenuate sunlight to
some degree so that only some of the grains are fully bleached,
the result will be a sharp, positively skewed distribution with a
mode near the younger, or leading, edge. They proposed using
the mean of the lowest 5% as the best estimate of D E and obtained reasonable results from some young fluvial sediments.
Lepper and McKeever (2002) studied this in more detail and
argued that the 5% was arbitrary and often underestimated the
true age of the deposit. They proposed a leading edge statistic
that attempted to define a normal distribution that would characterize a lower mode. They did this by fitting a Gaussian curve
to the leading edge of the distribution and taking the steepest
part of the slope (one standard deviation below the mode) as
the D E . The steeper the slope, the higher the precision. Others
have tried more sophisticated statistical approaches to isolate
the lowest, normal portion of the curve (Spencer et al 2003,
Roberts et al 2000).
Post-depositional turbation that mixes sediments of
different ages is a significant problem for sediment dating
(Bateman et al 2003). The best estimate of D E from the
top of a stratigraphic unit, where older grains have been
turbated upwards, might be the leading edge, but the best
estimate from the bottom of the unit, where younger grains
have been turbated downwards, might be the trailing edge
(Feathers 2003). Geological interpretation is clearly important
for understanding these distributions, but conversely these
distributions should also aid in geological interpretations,
especially where independent age information is available.
OSL dating of sediments could play a big role in understanding
site formation.
1503

J K Feathers

6.1. Anthropogenic sediments


Dating anthropogenic sediments is of obvious interest to
archaeologists. Direct dates for activities that produced the
sediment obviate reliance on associated artefacts or organic
material. Burbidge et al (2001) used OSL on quartz to date
three soils created by post-Neolithic agricultural activities as
well as intervening sand deposits on the Shetland Islands
and achieved a stratigraphically consistent sequence consistent
with other archaeological evidence. Our laboratory has applied
OSL to soils that occur below and within artificial earthen
mounds in order to date construction of these features (Bush
and Feathers 2003, Feathers 1997). Dating of palaeosols by
luminescence assumes that zeroing occurs by the recycling of
grains to the surface by turbation processes. The mechanism
for exposure is probably erosion from microrelief created by
processes such as tree-fall or animal burrowing. Once an
active soil is buried, surface exposure no longer occurs, so that
luminescence is dating the burial event, which for mounds is
the construction event. Presumably turbation is not successful
at bleaching all grains and the likelihood of exposure decreases
with depth in the soil. We tested this assumption by looking
at modern top soils from the top of mounds and palaeosols
from within and under mounds and found positively skewed
distributions of equivalent dose, the leading edge, mean and
range of which increased with depth (figure 7). Mound fill,
above and below the soils, produced older ages, suggesting
the mechanism of construction itself was not conducive to
bleaching.
Other kinds of archaeological construction datable by
luminescence are pits and canals, where dating of sediments
at their base provides a minimum date of abandonment. Lang
et al (1999) used IRSL to date fine-grained sediments that had
washed into prehistorically dug cellars from human-induced
erosion. By determining D E from different parts of the shinedown curves, they found no evidence of partial bleaching.
They also selected their samples to minimize the possibility
of admixture of older grains from the cellar walls. The derived
ages were consistent with other archaeological evidence.
Sanderson et al (2003) have used both OSL and IRSL to
date abandonment and use of ancient canals in the Cambodian
Mekong Delta. Cursory luminescence measurements on a
series of small samples collected from a stratigraphic profile
through one of the canals allowed separation of canal fill
from the underlying substrate and also identification of later
disturbance and other depositional features. More in-depth
study of selected samples provided dates consistent with
other archaeological data. Preliminary small-aliquot analysis
showed samples in the upper fill to be well bleached, but
basal samples gave evidence of substantial mixing with older
substrate (see also Spencer et al 2003). IRSL on feldspars from
the basal samples also suggested that fine-grained feldspars
may be better bleached in this depositional setting than coarser
grains or quartz.
Luminescence may also help in identifying anthropogenic
sediments. Roberts et al (2001b) dated 19 fine-grained samples from a thick loess deposit in northern China to determine
accumulation rates and to look for disturbance attributed to
agricultural practices. The authors used blue stimulation following a long IRSL exposure to minimize the OSL signal from
feldspars and therefore the potential of fading. A modified
1504

Figure 7. Histograms of equivalent dose distributions for segments


of a modern soil at the Watson Brake archaeological site in
northeastern Louisiana. The D E in the top 5 cm are clustered near
zero. Non-zero values become more common with depth as larger
percentages of grains have not been fully exposed through turbation.
The data are from single aliquots containing 1090 grains, with the
exception of the 510 cm segment where 200400 grain aliquots
were measured.

SAR protocol was adopted. The results showed that the deposition was quasi-continuous for at least 12 000 years. Fifteen
of the dates were in the correct stratigraphic order. The four
exceptions, confined to two time periods, gave ages that were
too high and were explained by agricultural activity that introduced unbleached grains into the deposits either by addition
of material to increase arability or through human reworking.
6.2. Site integrity
Two archaeological issues that have engendered a good
deal of debate are dates of colonization and correlation of
human activities with environmental events, such as the
extinction of large mammalian fauna. Claims for early
colonization or extinction correlations have been closely
scrutinized, particularly in terms of valid dating and site
integrity. Luminescence provides not only an independent
dating method but also, perhaps more importantly, an
assessment of the amount of mixing or disturbance in the
sediments. Interpretation of single-grain equivalent dose
distributions is already playing an important role in the
Australian debate. I have already mentioned Jinmium rock
shelter, where single-grain distributions revealed a mixed
deposit that earlier was not appreciated and led to severe

Use of luminescence dating in archaeology

overestimation of the age of human occupation there. In


contrast, distributions from Malakunanja II rock shelter were
consistent with a single bleaching date, indicating human
occupation 5060 ka (Roberts et al 1998b). These ages were
consistent with radiocarbon assays and agreed with earlier TL
results, which used multi-aliquot techniques and included a
signal less sensitive to light.
The age of the oldest known human skeletons in Australia,
from Lake Mungo, has been contested, with some OSL measurements, as well as U-series and ESR dating on the skeleton
remains themselves, giving an age as old as 60 ka (Thorne
et al 1999). Criticism of the OSL results has centred around
uncertainties in the environmental dose rate and the association of the dated sediments with the burials, but also over
the fact that the OSL was performed on multi-aliquots, which
could mask incomplete bleaching (Gillespie and Roberts 2000,
cf Grun et al 2000). Very recently a comprehensive OSL dating project involving 25 dates from three stratigraphic sections
bracketing two of the burials have constrained the dates of both
to 40 2 ka (Bowler et al 2003). Equivalent dose was determined from both multi-aliquots, using the slide method, and
from small (10 grains) single aliquots, using SAR. The mean
results were in agreement and the single-aliquot distributions
did not indicate mixing or partial bleaching.
Roberts and colleagues have also begun an ambitious
project to date the extinction of the Australian megafauna
(Roberts et al 2001a). OSL samples were secured from 28
palaeontological sites bearing on the issue and dated using
single aliquots containing typically less than 10 grains. For
samples with positively skewed equivalent dose distributions,
the youngest grains were taken to represent the equivalent dose,
using a minimum age model (Galbraith et al 1999). They found
no evidence for the megafauna extending later than about 46
ka and were therefore not able to rule out a human role in
extinction. One potentially younger occurrence of megafauna
at the controversial site of Cuddie Springs was questioned
because of evidence of sediment mixing, a conclusion strongly
disputed by the site excavators (Wroe and Field 2001).
OSL is also being applied to the contentious issue of
when people first settled in the Americas. Dating projects
are currently being carried out in Missouri, Virginia, South
Carolina and Brazil at relevant sites, but so far little has been
published, except for TL dates on lithics and OSL dates on
hearth sediments at an 11 ka site in the Amazon (Michab et al
1998).
Finally, OSL is being increasingly used to evaluate the age
and integrity of sites associated with early evidence of modern
human behaviour in sub-Saharan Africa (Henshilwood et al
2002, Feathers and Migliorini 2001).
6.3. Site formation
An added benefit of sediment dating is information on site
formation processes, understanding of which is essential
for not only assessing site integrity but to interpret spatial
patterning of artefacts.
Frederick et al (2002) used OSL dating to evaluate the
depositional process of an extensive sand mantle across east
Texas. The issue is important to archaeologists trying to
assess the significance of archaeological material found in

Figure 8. Histogram of equivalent dose from an aeolian sample


from Mustang Springs in west Texas, the same sample as illustrated
in figure 6. The distribution is negatively skewed, suggesting
extensive mixing in this deposit. The oldest grains, or trailing edge,
yield ages closest to that expected from independent evidence
(Feathers 2003).

these sands. One argument is that the sands are the product of
in situ weathering of Tertiary bedrock so that archaeological
materials found at depth can only be there because of
pedoturbation, their spatial patterning therefore having no
archaeological significance. An opposing view is that the
sands are aeolian Holocene deposits, which have preserved
archaeological spatial integrity. The authors reasoned that the
samples under the weathering model should have saturated
OSL signals but under the aeolian model should produce
signals that were well bleached on deposition and that give
Holocene ages in correct stratigraphic order. Single-aliquot
analysis was used and while the aeolian model in general
was supported, the distribution of equivalent doses did suggest
some post-depositional mixing on some samples, an issue the
authors are currently exploring (Bateman et al 2003).
Mixing from turbation or deflation in dune sediments
has long been apparent to archaeologists and geologists
(e.g. Mayer 2002). Figure 8 shows a negatively skewed
distribution of single-grain equivalent doses from a massive
aeolian deposit with little internal structure from the Mustang
Springs site in western Texas, suggesting extensive turbation.
The trailing edge was used to date the sample. Although
younger than a single radiocarbon determination from the same
level, this age is still preferable to the mean as an age for
deposition (Feathers 2003). The same study reported possible
turbation problems with other dune sediments, although other
aeolian depositions sampled, including some dunes, provided
reasonable dates. Sommerville et al (2001) sampled a number
of dune deposits in Scotland, associated with archaeological
sites, and found some variation in the extent of bleaching
during deposition, due in part to recycling from older dune
deposits, but nevertheless obtained ages that were broadly in
agreement with independent evidence. Additional work on
small aliquots (Spencer et al 2003) suggested there was some
mixing. A statistical analysis of Gaussian distributions using
F-ratios applied sequentially to equivalent dose values (from
low to high) suggested mixing from anthrogenic soils both
above and below the aeolian deposit and may allow estimation
of the time of abandonment between them.
1505

J K Feathers

Waterlain deposits from low-energy transport systems


have generally yielded successful results as well, but problems
have been reported.
Folz et al (2001) reported OSL
overestimation when compared with radiocarbon dates from a
Late Palaeolithic site located in Seine River deposits near Paris.
While discounting several possible reasons to account for the
overestimation, they did not have the ability to date single
grains. Possibly such resolution may have allowed detection of
the positive skewness noticed in equivalent dose distributions
of other fluvial deposits (Olley et al 1999, Lepper et al 2000)
and would imply a younger age.
Rock shelters have long been used for human occupation
and in many areas because of preservation reasons provide a
large fraction of the archaeological record. Rock shelters serve
as traps for sediment from a variety of sources, so the potential
for complex radiation environments, mixing of sediments and
partial bleaching is high. Most OSL attempts have targeted
aeolian-derived sediments in rock shelters (Schwarcz and Rink
2001, Feathers and Bush 2000) but single-aliquot and singlegrain dating should allow evaluation of colluvial sediments as
well.

7. Conclusion
Recent methodological developments in luminescence dating,
particularly single-aliquot and single-grain dating, have
increased accuracy and precision and have also extended the
application of luminescence beyond dating to depositional
processes and site formation.
Archaeologists are only
beginning to take advantage of these developments and have
indeed been slow at even appreciating the benefits of traditional
TL dating of pottery that has been available for years.
This paper has attempted to review these methodological
developments and to provide examples of applications that
demonstrate how questions, important to archaeological
method, can be addressed with luminescence. The direct
dating capabilities of pottery dating make luminescence an
ideal method for teasing apart mixed assemblages, determining
occupation durations and dating small sites that are otherwise
difficult to place in regional chronologies. The resolution of
single-aliquot and single-grain dating presents the opportunity
for archaeologists to date and identify anthropogenic deposits,
assess the stratigraphic integrity of sites and to determine site
formation processes. In many such applications, luminescence
should become the dating method of choice.

References
Agersnap Larsen N, Bulur E, Btter-Jensen L and McKeever
S W S 2000 Use of the LM-OSL technique for the
detection of partial bleaching in quartz Radiat. Meas. 32
41926
Aitken M J 1985 Thermoluminescence Dating (London: Academic)
Aitken M J 1998 An Introduction to Optical Dating (Oxford:
Oxford University Press)
Aitken M J and Xie J 1992 Optical dating using infrared diodes:
young samples Q. Sci. Rev. 11 14752
Akridge D G and Benoit P H 2001 Luminescence properties of chert
and some archaeological applications J. Archaeol. Sci. 28
14351
Anthony I M C, Sanderson D C W, Cook G T, Abernethy D and
Housley R A 2001 Dating a burnt mound from Kilmartin,
Argyll, Scotland Q. Sci. Rev. 20 9215

1506

Auclair M, Lamothe M and Huot S 2003 Measurement of anomalous


fading for feldspar IRSL using SAR Radiat. Meas. at press
Bailey R M 2001 Towards a general kinetic model for optically and
thermally stimulated luminescence of quartz Radiat. Meas. 33
1745
Bailey R M, Singarayer J, Ward S and Stokes S 2003 Identification
of partial bleaching using D E as a function of illumination time
Radiat. Meas. at press
Bailey R M, Smith B W and Rhodes E J 1997 Partial bleaching and
the decay form characteristics of quartz OSL Radiat. Meas. 27
12336
Banerjee D, Btter-Jensen L and Murray A S 2000 Retrospective
dosimetry: estimation of the dose to quartz using the
single-aliquot regenerative-dose protocol Appl. Radiat. Isot. 52
83144
Banerjee D, Murray A S, Btter-Jensen L and Lang A 2001
Equivalent dose estimation using a single aliquot of
polymineral fine grains Radiat. Meas. 33 7394
Barnett S M 2000a Luminescence dating pottery from later
prehistoric Britain Archaeometry 42 43157
Barnett S M 2000b Sampling pottery in luminescence dating studies
Radiat. Meas. 32 46772
Bateman M D, Singhvi A K and Frederick C D 2003 Investigations
into the potential effects of pedoturbation on luminescence
dating Q. Sci. Rev. 22 116976
Beck C 1994 Introduction Dating in Exposed and Surface Contexts
ed C Beck (Albuquerque: University of New Mexico Press)
pp 313
Btter-Jensen L, Duller G A T and Murray A S 2002 Advances in
luminescence measurements systems Radiat. Meas. 32 5238
Bowler J M, Johnson H, Olley J M, Prescott J R, Roberts R G,
Shawcross W and Spooner N A 2003 New ages for human
occupation and climatic change at Lake Mungo, Australia
Nature 421 83740
Bronk Ramsey C 1995 Radiocarbon calibration and analysis of
stratigraphy: the OxCal program Radiocarbon 37 425530
Bulur E, Btter-Jensen L and Murray A S 2000 Optically stimulated
luminescence from quartz measured using the linear
modulation technique Radiat. Meas. 32 40712
Burbidge C I, Batt C M, Barnett S M and Dockrill S J 2001 The
potential for dating the Old Scatness site, Shetland, by optically
stimulated luminescence Archaeometry 43 58996
Bush D A and Feathers J K 2003 Application of OSL dating to
anthrosols: a study of the distribution of D E of quartz through
surface and buried soil profiles from Archaic-aged earthen
mounds in the SE United States Q. Sci. Rev. 22 11539
Chiavari C, Martini M, Sibilia E and Vandini M 2001
Thermoluminescence (TL) characterization and dating
feasibility of ancient glass mosaic Q. Sci. Rev. 20 96772
Clarke M L, Rendell H M and Wintle A G 1999 Quality assurance
in luminescence dating Geomorphology 29 17385
Dean J S 1978 Independent dating in archaeological analysis Adv.
Archaeol. Method Theory 1 22355
Duller G A T 1995 Luminescence dating using single aliquots:
methods and applications Radiat. Meas. 24 21726
Duller G A T 1997 Behavioural studies of stimulated luminescence
from feldspars Radiat. Meas. 27 66394
Duller G A T, Btter-Jensen L and Murray A S 2000 Optical dating
of single sand-sized grains of quartz: sources of variability
Radiat. Meas. 32 4537
Dunnell R C and Feathers J K 1994 Thermoluminescence dating of
surficial archaeological material Dating in Exposed and
Surface Contexts ed C Beck (Albuquerque: University of New
Mexico Press) pp 11537
Dykeman D D, Towner R H and Feathers J K 2002 Correspondence
in tree-ring and thermoluminescence dating: a protohistoric
Navajo pilot study Am. Antiq. 67 14564
Fattahi M and Stokes S 2003 Absorbed dose evaluation in feldspar
using a single aliquot regenerative-dose (SAR) infrared
stimulated red luminescence protocol Radiat. Meas. at press
Feathers J K 1997 Luminescence dating of early mounds in
northeast Louisiana Q. Sci. Rev. 16 33340

Use of luminescence dating in archaeology

Feathers J K 2000a Luminescence dating and why it deserves wider


application Its About Time: A History of Archaeological
Dating in North America ed S Nash (Salt Lake City, UT:
The University of Utah Press) pp 15266
Feathers J K 2000b Date list 7: luminescence dates on prehistoric
and proto-historic pottery from the American Southwest
Ancient TL 18 5161
Feathers J K 2002 Luminescence dating in less than ideal
conditions: case studies from Klasies River Main
Site and Duinefontein, South Africa J. Archaeol. Sci. 29
17794
Feathers J K 2003 Luminescence dating of sediments from the
Southern High Plains Q. Sci. Rev. 22 103542
Feathers J K and Bush D A 2000 Luminescence dating of Middle
Stone Age deposits at Die Kelders J. Hum. Evolution 38
91119
Feathers J K and Migliorini E 2001 Luminescence dating at
Katandaa reassessment Q. Sci. Rev. 20 9616
Folz E, Bodu P, Bonte P, Joron J-L, Mercier N and Reyss J-L 2001
OSL dating of fluvial quartz from Le Closeau: a Late
Prehistoric site near Paris Q. Sci. Rev. 20 92733
Franklin A D and Hornyak W F 1990 Isolation of the rapidly
bleaching peak in quartz TL glow curves Ancient TL
10 16
Franklin A D, Hornyak W F and Dickerson W 1992 TL estimation
of paleodose of dune-sand quartz Q. Sci. Rev. 11 758
Franklin A D, Prescott J R and Scholefield R B 1995 The
mechanism of thermoluminescence in an Australian
sedimentary quartz J. Lumin. 63 31726
Frederick C D, Bateman M D and Rogers R 2002 Evidence for
eolian deposition in the sandy uplands of east Texas and the
implications for archaeological site integrity Geoarchaeology
17 191217
Galbraith R F, Roberts R G, Laslett G M, Yoshida H and Olley J M
1999 Optical dating of single and multiple grains of quartz from
Jinmium rock shelter, northern Australia: Part I, experimental
design and statistical models Archaeometry 41 33964
Gallivan M D 2002 Measuring sedentariness and settlement
population: accumulations research in the Middle Atlantic
region Am. Antiq. 67 53553
Gautier A 2001 Luminescence dating of archeometallurgical slag:
use of the SAR technique for determination of the burial dose
Q. Sci. Rev. 20 97380
Gemmell A M D 1997 Fluctuations in the thermoluminescence
signal of suspended sediment in an alpine glacial meltwater
stream Q. Sci. Rev. 16 28190
Gillespie R and Roberts R G 2000 On the reliability of age estimates
for human remains at Lake Mungo J. Hum. Evolution 38
72732
Godfrey-Smith D I, Deal M and Kunelius I 1997
Thermoluminescence dating of St Croix ceramics: chronology
building in southwestern Nova Scotia Geoarchaeology 12
25173
Greilich S, Glasmacher U A and Wagner G A 2002 Spatially
resolved detection of luminescence: a unique tool for
archaeochronometry Naturwissenschaften 89 3715
Grun R, Spooner N A, Thorne A, Mortimer G, Simpson J J,
McCulloch M T, Taylor L and Curnoe D 2000 Age of
the Lake Mungo 3 skeleton, reply to Bowler and Magee
and to Gillespie and Roberts J. Hum. Evolution 38
73341
Guibert P, Schvoerer M, Etcheverry M P, Szepertyski B and
Ney C 1994 IXth millennium B C ceramics from Niger:
detection of a U-series disequilibrium and TL dating Q. Sci.
Rev. 13 55561
Guibert P, Vartanian E, Roque C, Schvoerer M and Bechtel F 2001
Luminescence dating of burnt materials: effects of preheat
treatment on OSL and consequences for dating procedures
Radiat. Meas. 33 43944
Habermann J, Schilles T, Kalchgruber R and Wagner G A 2000
Steps towards surface dating using luminescence Radiat. Meas.
32 84751

Hashimoto T, Komatsu Y, Hong D G and Uezu Y 2001


Determination of radionuclides in small pieces of
archaeological sample and its application to TL-dating Radiat.
Meas. 33 95101
Haustein M, Krbetschek M R, Trautmann T, Roewer G and
Stolz W 2001 A luminescence study for dating
archaeometallurgical slags Q. Sci. Rev. 20 9815
Henshilwood C S, dErrico F, Yates R, Jacobs Z, Tribolo C,
Duller G A T, Mercier N, Sealy J C, Valladas H, Watts I and
Wintle A G 2002 Emergence of modern human behaviour:
Middle Stone age engravings from South Africa Science 295
127880
Herbert J M, Feathers J K and Cordell A 2002 Building ceramic
chronologies with thermoluminescence dating: a case study
from the Carolina Sandhills Southeastern Archaeol. 21
92108
Holt J Z and Feathers J K 2003 Dating the Middle to Late Woodland
transition in the Illinois valley: radiocarbon and TL dates from
the BaehrGust site Midcontinental J. Archaeol. at press
Hong D G, Yi S B, Galloway R B, Tsuboi T and Hashimoto T 2001
Optical dating of archaeological samples using a single aliquot
of quartz stimulated by blue light J. Radioanal. Nucl. Chem.
247 17984
Huntley D J, Godfrey-Smith D I and Thewalt M L W 1985 Optical
dating of sediments Nature 313 1057
Huntley D J, Hutton J T and Prescott J R 1993 The stranded
beach-dune sequence of southeast South Australia: a test of
luminescence dating, 0-800 ka Q. Sci. Rev. 12 120
Huntley D J and Lamothe M 2001 Ubiquity of anomalous fading in
K-feldspars, and measurement and correction for it in optical
dating Can. J. Earth Sci. 38 1093106
Huntley D J, Short M A and Dunphy K 1996 Deep traps in quartz
and their use for optical dating Can. J. Phys. 74 8191
Hutt G, Jaek I and Tchonka J 1988 Optical dating: K-feldspars
optical response stimulation spectra Q. Sci. Rev. 7 3816
Kojo Y 1991 The reliability of thermoluminescence dating: a pilot
experiment Geoarchaeology 6 36774
Krbetschek M R, Gotze J, Dietrich A and Trautmann T 1997
Spectral information from minerals relevant for luminescence
dating Radiat. Meas. 27 695748
Lamothe M and Auclair M 2000 The fadia method: a new approach
in luminescence dating using the analysis of single feldspar
grains Radiat. Meas. 32 4338
Lang A, Kadereit A, Behrends R-H and Wagner G A 1999 Optical
dating of anthropogenic sediments at the archaeological site of
Herrenbrunnenbuckel, Bretten-Bauerbach (Germany)
Archaeometry 41 397411
Lepper K, Agersnap Larsen N and McKeever S W S 2000
Equivalent dose distribution analysis of Holocene eolian and
fluvial quartz sands from central Oklahoma Radiat. Meas. 32
6038
Lepper K and McKeever S W S 2002 An objective methodology for
dose distribution analysis Radiat. Prot. Dosim. 101 34952
Li S-H 1994 Optical dating: insufficiently bleached sediments
Radiat. Meas. 23 5637
Mayer J H 2002 Evaluating natural site formation processes in eolian
dune sands: a case study from the Krmpotich Folsom site,
Killpecker Dunes, Wyoming J. Archaeol. Sci. 29 1199211
McCoy D G, Prescott J R and Nation R J 2000 Some aspects of
single-grain luminescence dating Radiat. Meas. 32 85964
McFee C J and Tite M S 1998 Luminescence dating of
sedimentsthe detection of high equivalent dose grains using
an imaging photon detector Archaeometry 40 15368
McKeever S W S and Chen R 1997 Luminescence models Radiat.
Meas. 27 62561
Mercier N and Valladas H 1994 Thermoluminescence dates for the
Paleolithic Levant Late Quaternary Chronology and
Palaeoclimate of the Eastern Mediterranean ed O Bar-Yosef
and R Kra (Tuscon: University of Arizona) pp 1320
Mercier N, Valladas H and Froget L 1999 Thermoluminescence
dating of a middle Palaeolithic occupation at Sodmein Cave,
Red Sea Mountains (Egypt) J. Archaeol. Sci. 26 133945

1507

J K Feathers

Mercier N, Valladas H, Froget L, Joron J-L and Ronen A 2000


Datation par thermoluminescence de la base du gisement
paleolithique de Tabun (Mont Carmel, Israel) C. R. Acad. Sci.,
Paris II 330 7318
Mercier N, Valladas H, Joron J-L, Schiegl S, Bar-Yosef O and
Weiner S 1995a Thermoluminescence dating and the problem
of geochemical evolution of sedimentsa case study: the
Mousterian levels at Hayonim (1994) Israel J. Chem. 35
13741
Mercier N, Valladas H and Valladas G 1995b Flint
thermoluminescence dates from the CFR Laboratory at Gif:
contributions to the study of the chronology of the Middle
Palaeolithic Q. Sci. Rev. 14 35164
Michab M, Feathers J K, Joron J-L, Mercier N, Selos M, Valladas H,
Valladas G, Reyss J-L and Roosevelt A C 1998 Luminescence
dates for the paleoindian site of Pedra Pintada, Brazil Q. Sci.
Rev. 17 10416
Morganstein M E, Luo S, Ku T-L and Feathers J K 2003 Uranium
series weathering rind and luminescence dating of volcanic
lithic artifacts Archaeometry at press
Murray A S and Clemmensen L B 2001 Luminescence dating of
Holocene aeolian sand movement, Thy, Denmark Q. Sci. Rev.
20 7514
Murray A S and Mejdahl V 1999 Comparison of regenerative-dose
single aliquot and multi-aliquot (SARA) protocols using heated
quartz from archaeological sites Quaternary Geochronology
(Q. Sci. Rev.) 18 2239
Murray A S and Roberts R G 1998 Measurements of the equivalent
dose in quartz using a regenerative-dose single-aliquot protocol
Radiat. Meas. 29 50315
Murray A S and Wintle A G 2000 Luminescence dating of quartz
using an improved single-aliquot regenerative-dose protocol
Radiat. Meas. 32 5773
Murray A S and Wintle A G 2003 The single-aliquot
regenerative-dose protocol: potential for improvements in
reliability Radiat. Meas. at press
Nash S E 1997 A cutting-date estimation technique for ponderosa
pine and Douglas fir wood specimens Am. Antiq. 62 26072
G and Yurdatapan E 2000 Optically stimulated luminescence
Oke
dating of pottery from Turkey Talanta 53 11519
Olley J, Caitcheon G and Roberts R G 1999 The origin of dose
distributions in fluvial sediments, and the prospect of dating
single grains from fluvial deposits using optically stimulated
luminescence Radiat. Meas. 30 20717
Olley J M, Murray A S and Roberts R G 1996 The effects of
disequilibria in the uranium and thorium decay chains
on burial dose rates in fluvial sediments Q. Sci. Rev. 15
75160
Porat N, Chazan M, Schwarcz H and Horwitz L K 2002 Timing of
the lower to middle Paleolithic boundary: new dates from the
Levant J. Hum. Evolution 43 10722
Prescott J R, Huntley D J and Hutton J T 1993 Estimation of
equivalent dose in thermoluminescence datingthe Australian
slide method Ancient TL 11 15
Prescott J R, Robertson G B and Green J C 1982
Thermoluminescence dating of Pacific Island pottery:
successes and failures Archaeol. Oceania 17 1427
Readhead M L 1987 Thermoluminescence dose rate data
and dating equations for the case of disequilibrium in the
uranium decay series Nucl. Tracks Radiat. Meas. 13 197207
Readhead M L 1988 Thermoluminescence dating of quartz in
aeolian sediments from southeastern Australia Q. Sci. Rev. 7
25764
Richter D, Waiblinger J, Rink W J and Wagner G A 2000
Thermoluminescence, electron spin resonance and 14 C-dating
of the late middle and early upper Palaeolithic site of
Geienklosterle Cave in southern Germany J. Archaeol. Sci. 27
7189
Roberts H M, Wintle A G, Maher B A and Hu M 2001b Holocene
sediment-accumulation rates in the western Loess Plateau,
China, and a 2500-year record of agricultural activity, revealed
by OSL dating The Holocene 11 47783

1508

Roberts R G 1997 Luminescence dating in archaeology: from


origins to optical Radiat. Meas. 27 81992
Roberts R G, Bird M, Olley J, Galbraith R, Lawson E, Laslett G,
Yoshida H, Jones R, Fullagar R, Jacobsen G and Hua Q 1998a
Optical and radiocarbon dating at Jinmium rock shelter in
northern Australia Nature 393 35862
Roberts R G, Flannery T F, Ayliffe L K, Yoshida H, Olley J M,
Prideaux G J, Laslett G M, Baynes A, Smith M A, Jones R and
Smith B L 2001a New ages for the last Australian megafauna:
continent-wide extinction about 46 000 years ago Science 292
188892
Roberts R G, Galbraith R F, Olley J M, Yoshida H and Laslett G M
1999 Optical dating of single and multiple grains of quartz
from Jinmium rock shelter, northern Australia. Part II: results
and implications Archaeometry 41 36595
Roberts R G, Galbraith R F, Yoshida H, Laslett G M and Olley J M
2000 Distinguishing dose populations in sediment mixtures: a
test of single-grain optical dating procedures using mixtures of
laboratory-dosed quartz Radiat. Meas. 32 45965
Roberts R G, Yoshida H, Galbraith R F, Laslett G M, Jones R and
Smith M 1998b Single-aliquot and single-grain optical dating
confirm thermoluminescence age estimates at Malakunanja II
rock shelter in northern Australia Ancient TL 16 1924
Sampson C G, Bailiff I and Barnett S 1997 Thermoluminescence
dates from Later Stone Age pottery on surface sites in the
Upper Karoo S. Afr. Archaeol. Bull. 52 3842
Sampson E H, Fleming S J and Bray W 1972 Thermoluminescent
dating of Colombian pottery in the Yotoco style Archaeometry
14 1218
Sanderson D C W, Bishop P and Stark M 2003 Luminescence
dating of anthropogenically re-set canal sediments from
Angkor Borei, Mekong Delta, Cambodia Q. Sci. Rev.
22 111121
Schiffer M B 1975 Archaeology as behavioural science Am.
Anthropol. 77 83648
Schiffer M B 1986 Radiocarbon dating and the old wood problem:
the case of the Hohokam chronology J. Archaeol. Sci. 13 1330
Schwarcz H P 1994 Current challenges to ESR dating Q. Sci. Rev.
13 6015
Schwarcz H P and Rink W J 2001 Dating methods for sediments of
caves and rockshelters with examples from the Mediterranean
region Geoarchaeology 16 35571
Singarayer J S and Bailey R M 2003 Further investigations of the
quartz OSL components using the linear modulated technique
Radiat. Meas. at press
Sommerville A A, Sanderson D C W, Hansom J D and Housley R A
2001 Luminescence dating of aeolian sands from
archaeological sites in northern Britain: a preliminary study Q.
Sci. Rev. 20 91319
Spencer J Q, Sanderson D C W, Deckers K and Sommerville A A
2003 Assessing mixed dose distributions in young sediments
identified using small aliquots and a simple two-step SAR
procedure: the F-statistic as a diagnostic tool Radiat. Meas. at
press
Spooner N A 1994a The anomalous fading of infrared-stimulated
luminescence from feldspars Radiat. Meas. 23 62532
Spooner N A 1994b On the optical dating signal from quartz Radiat.
Meas. 23 593600
Stein J K 1987 Deposit for archaeologists Adv. Archaeol. Method
Theory 2 33793
Theocaris P S, Liritzis I and Galloway R B 1997 Dating of two
Hellenic pyramids by a novel application of
thermoluminescence J. Archaeol. Sci. 24 399405
Thorne A, Grun R, Mortimer G, Spooner N A, Simpson J J,
McCulloch M, Taylor L and Curnoe D 1999 Australias oldest
human remains: age of the Lake Mungo 3 skeleton J. Hum.
Evolution 36 591612
Valladas H 1992 Thermoluminescence dating of flint Q. Sci. Rev. 11
15
Valladas G and Gillot P Y 1978 Dating of the Olby lava flow using
heated quartz pebbles: some problems PACT 2 1419

Use of luminescence dating in archaeology

Valladas H, Mercier N, Froget L, Joron J-L and Reyss J-L 1998 GIF
laboratory dates for Middle Paleolithic Levant Neanderthals
and Modern Humans in Western Asia ed T Akazawa,
K Aoki and O Bar-Yosef (New York: Plenum) pp 6975
Valladas H, Mercier N, Froget L, Joron J-L, Reyss J-L and
Aubry T 2001 TL dating of Upper Palaeolithic sites in the Coa
Valley (Portugal) Q. Sci. Rev. 20 93943
Varien M D and Mills B J 1997 Accumulations research: problems
and prospects J. Archaeol. Method Theory 4 14191
Visocekas R, Spooner N A, Zink A and Blanc P 1994 Tunnel
afterglow, fading and infrared emission in thermoluminescence
of feldspars Radiat. Meas. 23 37785
Vogel J C, Wintle A G and Woodborne S M 1999 Luminescence
dating of coastal sands: overcoming changes in environmental
dose rate J. Archaeol. Sci. 26 72933

Wallinga J, Murray A and Wintle A 2000 The single-aliquot


regenerative-dose (SAR) protocol applied to coarse-grain
feldspar Radiat. Meas. 32 52934
Whittle E H and Arnaud J M 1975 Thermoluminescent dating of
Neolithic and Chalcolithic pottery from sites in central
Portugal Archaeometry 17 524
Wilhelmsen K and Miles T 2003 The application of
thermoluminescence dating to catastrophically heated chert
artifacts New Methods of Lithic Analysis ed M Shott and
N Maloney (London: Institute of Archaeology, University
College) at press
Wroe S and Field J 2001 Mystery of megafaunal extinction remains
Australas. Sci. 22 215
Zink A J C and Visocekas R 1997 Datability of sanidine feldspars
using the near-infrared TL emission Radiat. Meas. 27 25161

1509

Potrebbero piacerti anche