13.11.2013 Views

Luminescence chronologies for coastal and marine sediments

Luminescence chronologies for coastal and marine sediments

Luminescence chronologies for coastal and marine sediments

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Luminescence</strong> <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong><br />

ZENOBIA JACOBS<br />

BOREAS<br />

Jacobs, Z. 2008 (November): <strong>Luminescence</strong> <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong>. Boreas, Vol, 37, pp.<br />

508–535. 10.1111/j.1502-3885.2008.00054.x. ISSN 0300-9483.<br />

<strong>Luminescence</strong> dating techniques have been applied to a range of different types of <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> deposits,<br />

involving a variety of depositional processes <strong>and</strong> environments, as a means of addressing an assortment of local,<br />

regional <strong>and</strong> global themes. This review provides a brief historical background of 30 years of luminescence dating<br />

applied to <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong>, followed by a discussion of some of the luminescence dating highlights <strong>for</strong><br />

nine major regions of the world. A literature review of 200 readily accessible peer-reviewed journal publications is<br />

summarised in a table <strong>and</strong> studies that have made some methodological advances when applying luminescence<br />

techniques to <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> deposits are discussed in detail. Where possible, a specific theme is explored <strong>for</strong><br />

each region. Potential pitfalls <strong>and</strong> future prospects of luminescence dating applied to <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong><br />

are also considered.<br />

Zenobia Jacobs (e-mail: zenobia@uow.edu.au), GeoQuEST Research Centre, School of Earth <strong>and</strong> Environmental<br />

Sciences, University of Wollongong, Wollongong, NSW 2522, Australia; received 18th March 2008, accepted 18th<br />

July 2008.<br />

Coastal areas around the world represent one of the<br />

most populated, extensive, diverse <strong>and</strong> dynamic l<strong>and</strong>scapes<br />

on earth. The <strong>coastal</strong> environment is characterized<br />

by a large range of l<strong>and</strong><strong>for</strong>ms <strong>and</strong> erosional <strong>and</strong><br />

depositional features that are in a process of constant<br />

change. Natural processes such as eustatic <strong>and</strong> isostatic<br />

sea-level change, tectonism, waves, wind <strong>and</strong> various<br />

weather phenomena have caused, <strong>and</strong> are continuing to<br />

cause, the erosion, accretion <strong>and</strong> reshaping of coasts.<br />

The coastlines of the world’s continents measure several<br />

hundreds of thous<strong>and</strong>s of kilometres in total length<br />

<strong>and</strong> a large proportion of human populations live in<br />

areas along or near the coast. As such, underst<strong>and</strong>ing<br />

the <strong>for</strong>mation of our coasts, <strong>and</strong> assessing the past,<br />

present <strong>and</strong> future impact of change on the evolution of<br />

our coasts, is there<strong>for</strong>e important <strong>and</strong> necessitates the<br />

use of a reliable <strong>and</strong> useful chronological tool.<br />

The potential <strong>and</strong> desirability of luminescence dating<br />

methods to determine the magnitude <strong>and</strong> timing of<br />

events recorded in <strong>coastal</strong> deposits have long been recognized.<br />

Coastal <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> were used in<br />

both the original development <strong>and</strong> validation of thermoluminescence<br />

(TL) dating of unheated <strong>sediments</strong><br />

<strong>and</strong> optically stimulated luminescence (OSL) dating of<br />

both feldspar <strong>and</strong> quartz grains (Huntley et al.<br />

1993a, b). Among the main attractions of luminescence<br />

dating are: (1) the fact that it can be applied to ubiquitously<br />

occurring materials (quartz <strong>and</strong> feldspar grains),<br />

(2) it dates directly the material <strong>and</strong> depositional event<br />

of interest, (3) its applicability over a long time range,<br />

<strong>and</strong> (4) its reporting of ages in calendar years. Although<br />

luminescence dating has been applied to <strong>coastal</strong> <strong>and</strong><br />

<strong>marine</strong> <strong>sediments</strong> since 1979, its routine use on a large<br />

scale has not yet been achieved. It has been shown to be<br />

useful in regions where there is a lack of suitable materials<br />

<strong>for</strong> other methods, e.g. corals <strong>for</strong> uranium-series<br />

(U-series) dating or organic materials <strong>for</strong> radiocarbon<br />

( 14 C) dating. Although it has been instrumental in determining<br />

the age of many <strong>coastal</strong> features around the<br />

world, the lack of precision, by comparison to 14 C <strong>and</strong><br />

U-series dating, has prevented it from making breakthrough<br />

contributions to issues such as the construction<br />

of a sea-level curve derived from far-field sites located<br />

on passive margins, even though it has the potential to<br />

do so. Recent progress in both methodology <strong>and</strong> technology,<br />

however, now allows increased precision <strong>and</strong><br />

accuracy. New methodological approaches also allow<br />

demonstration <strong>and</strong> explicit testing of the reliability of<br />

the ages, making luminescence dating a very attractive<br />

<strong>and</strong> increasingly used approach.<br />

A review of luminescence dating applied to such a<br />

diverse environment is there<strong>for</strong>e not just challenging<br />

but necessarily broad. The aim of this review is to provide<br />

a summary of such studies, indicating their locations,<br />

the <strong>coastal</strong> features being dated, the time epoch<br />

of interest <strong>and</strong> the broad geological or geomorphological<br />

theme each study aims to address. By bringing together<br />

the studies, several interesting trends within<br />

regions can be observed <strong>and</strong> these will be discussed. In<br />

addition, the aim is also to highlight potential pitfalls<br />

<strong>and</strong> future prospects of luminescence dating applied to<br />

young <strong>and</strong> old <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong>.<br />

Thirty years of luminescence dating applied to<br />

<strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong><br />

A comprehensive literature survey of luminescence<br />

dating of <strong>sediments</strong> over the past 30 years resulted in<br />

196 readily available peer-reviewed publications in<br />

DOI 10.1111/j.1502-3885.2008.00054.x r 2008 The Author, Journal compilation r 2008 The Boreas Collegium


BOREAS <strong>Luminescence</strong> <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> 509<br />

which luminescence-dating techniques were applied to<br />

<strong>coastal</strong> <strong>and</strong> <strong>marine</strong> deposits. From this survey, a total<br />

of 61 studies employed thermoluminescence (TL) dating<br />

of quartz or feldspars, 32 studies employed single<strong>and</strong><br />

multiple-aliquot infrared-stimulated luminescence<br />

(IRSL) dating of potassium-rich feldspar grains, 26<br />

studies used a combination of TL <strong>and</strong> IRSL or optically<br />

stimulated luminescence (OSL) dating of quartz<br />

grains <strong>and</strong> 69 studies employed OSL dating of quartz<br />

grains. Among these are studies that contributed significantly<br />

to the development, improvement <strong>and</strong> validation<br />

of both TL <strong>and</strong> OSL methods since the very<br />

beginning.<br />

Marine <strong>sediments</strong> extracted from deep-sea cores<br />

drilled in the Antarctic <strong>and</strong> North Pacific oceans were<br />

used to develop the first suitable procedures (the ‘total<br />

bleach’ <strong>and</strong> ‘partial bleach’ procedures) <strong>for</strong> TL dating<br />

of unheated <strong>sediments</strong> (Wintle & Huntley 1979a, 1980,<br />

1982). Initially, it was thought that TL was emitted<br />

from carbonates (Bothner & Johnson 1969) <strong>and</strong> radiolaria<br />

(Huntley & Johnson 1976) found within these<br />

<strong>sediments</strong>, but Wintle & Huntley (1979a) demonstrated<br />

that it was, in fact, primarily the detrital, fine silt feldspar<br />

grains adhering to the radiolaria that emitted the<br />

TL that increased with depth below the ocean floor.<br />

A Pleistocene <strong>coastal</strong> dune s<strong>and</strong> from southeast<br />

South Australia <strong>and</strong> a modern beach s<strong>and</strong> from the<br />

Canadian Pacific coast were likewise used to develop<br />

<strong>and</strong> validate the first suitable optical dating technique<br />

(Huntley et al. 1985a). Quartz grains were stimulated<br />

with green (514 nm) light from an argon-ion laser <strong>and</strong><br />

were preheated to 2501C. These measurements resulted<br />

in an age consistent with independent age estimates <strong>for</strong><br />

the dune s<strong>and</strong>, <strong>and</strong> a higher than expected, but significantly<br />

smaller, equivalent dose (D e ) than previously<br />

obtained using TL dating <strong>for</strong> the modern beach s<strong>and</strong>.<br />

The latter indicated that even though much shorter exposures<br />

to sunlight are required to empty the trapped<br />

charge populations used in optical dating, partial<br />

bleaching may still be a problem in more complex<br />

depositional environments.<br />

Even though two breakthrough advances were made<br />

using <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> to validate the<br />

techniques of TL <strong>and</strong> OSL, very few additional studies<br />

applying these techniques to <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> deposits<br />

were published during the 1980s. This review identified<br />

only five additional articles, all using TL dating,<br />

<strong>and</strong> those represented a range of locations from <strong>marine</strong><br />

<strong>sediments</strong> extracted from a core in the Arctic Sea<br />

(Berger et al. 1984), Hudson Bay in Canada (Forman<br />

et al. 1987), to Spain <strong>and</strong> Morocco in the Mediterranean<br />

(Brückner 1986), the ‘old red s<strong>and</strong>s’ in Sri Lanka<br />

(Singhvi et al. 1986), the Spencer Gulf (Smith et al.<br />

1982) <strong>and</strong> the prominent str<strong>and</strong>ed beach dune sequence<br />

(Huntley et al. 1985b), both in South Australia.<br />

The 1990s was a more productive period, both in<br />

terms of application <strong>and</strong> methodological developments,<br />

resulting in the publication of 58 articles. These were<br />

primarily dominated by TL dating (n = 41), particularly<br />

TL dating applied to a range of l<strong>and</strong><strong>for</strong>ms along the<br />

Australian coast (n = 33). The most significant methodological<br />

advances were the development of the ‘selective<br />

bleach’ (e.g. Prescott & Fox 1990; Prescott &<br />

Mojarrabi 1993) <strong>and</strong> ‘Australian slide’ (Prescott et al.<br />

1993) TL procedures, as well as the first single-aliquot<br />

optical dating procedure (Duller 1991). Both TL procedures<br />

were developed <strong>and</strong> tested extensively on the<br />

800 000-year-old <strong>coastal</strong> dune sequence in South Australia<br />

(e.g. Huntley et al. 1993a, 1994a). The ‘selective<br />

bleach’ involves D e estimation based on only the most<br />

light sensitive TL peak at 3251C, by separating it from<br />

the other TL peaks through selective optical bleaching<br />

using green <strong>and</strong> longer wavelength light. The ‘Australian<br />

slide’ involved the use of two sets of aliquots to which<br />

additive dose <strong>and</strong> regenerative dose methods were applied<br />

<strong>and</strong> the data then compared <strong>and</strong>, if appropriate,<br />

combined. The advantage of this method is the detection<br />

of any sensitivity change that may have occurred during<br />

the initial bleach of the regenerated aliquots, as well as<br />

overcoming problems associated with extrapolation of<br />

additive dose growth curves <strong>for</strong> older samples.<br />

These initial TL <strong>and</strong> optical dating protocols, however,<br />

both required that many aliquots (typically<br />

20–50), each consisting of many thous<strong>and</strong>s of sediment<br />

grains, be used to construct a sample’s dose response.<br />

With these methods, much of the uncertainty in the D e<br />

results from the inherent variability of the sensitivities<br />

of the many thous<strong>and</strong>s of grains that make up each<br />

aliquot. Furthermore, it is difficult using multiplealiquot<br />

techniques to produce accurate ages in environments<br />

where not all the sediment grains have received<br />

sufficient sunlight exposure prior to final burial (e.g.<br />

tsunami deposits or storm deposits). Huntley et al.<br />

(1985a) suggested that an important advantage of optical<br />

dating over TL dating is that only a single aliquot<br />

of sediment is required to construct a sample’s growth<br />

curve. This would allow replicate measurements of D e<br />

to be generated <strong>for</strong> the same sample. These provide an<br />

internal check on the reproducibility of results, thereby<br />

facilitating the recognition of sample contamination,<br />

partial bleaching <strong>and</strong> other problems that may need to<br />

be addressed be<strong>for</strong>e final age determination. Using<br />

<strong>coastal</strong> dune s<strong>and</strong>s from the west coast of the North<br />

Isl<strong>and</strong> of New Zeal<strong>and</strong>, Duller (1991, 1994, 1995) developed<br />

the first single-aliquot procedure <strong>for</strong> dating<br />

potassium-rich feldspar grains. Initially, only a small<br />

number of studies employed single-aliquot procedures<br />

(e.g. Duller 1996; Wintle et al. 1998; Clarke et al. 1999),<br />

but these studies stimulated further investigations into<br />

the use of single aliquots of both feldspar <strong>and</strong> quartz.<br />

The benefits of single aliquots, the improved control on<br />

accuracy <strong>and</strong> precision, <strong>and</strong> its applicability to a wider<br />

range of <strong>sediments</strong> <strong>and</strong> depositional environments<br />

found within the <strong>coastal</strong> environment were realized.


510 Zenobia Jacobs BOREAS<br />

This change is reflected in studies carried out since<br />

2000, where only 9 studies used TL dating, 12 studies<br />

used TL <strong>and</strong> IRSL/OSL, 10 studies used multiple-aliquot<br />

additive dose (MAAD) IRSL <strong>and</strong> the remaining<br />

70 studies employed single-aliquot OSL or IRSL. Although<br />

development of the single-aliquot regenerativedose<br />

(SAR) procedure of Murray & Wintle (2000) did<br />

not involve the use of any <strong>coastal</strong> or <strong>marine</strong> <strong>sediments</strong>,<br />

the success of the method with regard to increased precision<br />

<strong>and</strong> accuracy led to a significant increase in the<br />

number of published studies applied to <strong>coastal</strong> <strong>and</strong><br />

<strong>marine</strong> <strong>sediments</strong>. The first applications using SAR<br />

appeared in the 2001 LED proceedings published in<br />

Quaternary Science Reviews. Since then, a third of all<br />

published studies involving <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong><br />

have used the SAR procedure, many of which<br />

focus on obtaining <strong>chronologies</strong> <strong>for</strong> very young <strong>coastal</strong><br />

<strong>sediments</strong> (late-Holocene to modern), where calibration<br />

of 14 C ages is unreliable (e.g. Brooke et al. 2008c).<br />

What, when <strong>and</strong> where?<br />

Details of the 196 studies are summarized in Table 1.<br />

For each study, its geographic location, the type of<br />

l<strong>and</strong><strong>for</strong>m, feature or sediment dated, relevant epoch<br />

(Pleistocene or Holocene), luminescence technique used<br />

(OSL, IRSL or TL), <strong>and</strong> the theme the study aims to<br />

address are provided, along with the appropriate reference<br />

source.<br />

Among these studies, very many different types of<br />

features have been dated. The main features include<br />

beaches, beach ridges, raised beaches, beachrocks,<br />

<strong>marine</strong> terraces, aeolianites, <strong>for</strong>edunes, back-barrier<br />

dunes or dune cordons, palaeosols, tsunami deposits,<br />

storm deposits <strong>and</strong>, to a lesser extent, estuarine <strong>and</strong><br />

tidal-flat <strong>sediments</strong>.<br />

An equally diverse number of issues have been addressed<br />

through the application of luminescence dating<br />

to <strong>coastal</strong> deposits. These may be of very local or regional<br />

concern or are relevant to underst<strong>and</strong>ing global<br />

topics. Issues include, among others: (1) relative sea-level<br />

change, (2) identification of regionally correlative<br />

periods of increased aeolian activity in <strong>coastal</strong> zones,<br />

(3) regional neotectonic changes <strong>and</strong> the determination<br />

of <strong>coastal</strong> uplift <strong>and</strong> subsidence rates, (4) development<br />

of spits <strong>and</strong> barrier isl<strong>and</strong>s, (5) timing <strong>and</strong> frequency of<br />

increased storminess, (6) timing <strong>and</strong> recurrence intervals<br />

of tsunamis, (7) reconstruction of the spatial <strong>and</strong><br />

temporal development of local coastlines, (8) effects of<br />

climate change <strong>and</strong> rising sea levels on past, current <strong>and</strong><br />

future human populations to improve <strong>coastal</strong> management<br />

planning, <strong>and</strong> (9) the possible effects of global<br />

warming on ‘soft’ coasts. Many of these topics are specific<br />

to certain regions of the world. Table 1 is divided<br />

into nine geographical regions, namely: 1) Africa, 2)<br />

Australia, 3) Remote Oceania (including New Zeal<strong>and</strong><br />

<strong>and</strong> the Pacific Isl<strong>and</strong>s), 4) the Americas, 5) Asia, 6)<br />

Middle East, 7) Mediterranean, 8) northwestern Europe,<br />

<strong>and</strong> 9) the oceans <strong>and</strong> seas. The locations are<br />

plotted on a map of the world in Fig. 1, where the studies<br />

are distinguished by the luminescence technique<br />

used. Table 1 <strong>and</strong> Fig. 1 show that most dating studies<br />

are clustered around the coasts of Australia <strong>and</strong> New<br />

Zeal<strong>and</strong>, northwestern Europe <strong>and</strong> the Mediterranean,<br />

South Africa <strong>and</strong> North America. This is in stark contrast<br />

to those many hundreds of thous<strong>and</strong>s of kilometres<br />

of coastline that remain undated by<br />

luminescence methods, notably large parts of Asia,<br />

Canada, South America <strong>and</strong> Africa. Some of the major<br />

studies within each of these regions <strong>and</strong> their relevance<br />

to the general themes are discussed below. Where patterns<br />

emerge within a region, these are discussed.<br />

Africa<br />

There has only been a single study conducted on <strong>coastal</strong><br />

<strong>sediments</strong> outside southern Africa <strong>and</strong> the Mediterranean<br />

coast of North Africa. This was on the very substantial<br />

deposits bordering the Atlantic Ocean near<br />

Casablanca, Morocco, spanning the last one million<br />

years (Rhodes et al. 2006). The sequence is constructed<br />

from deposits at a number of different sites (Reddad<br />

Ben Ali, Oudad J’mel, Sidi Abderhamane <strong>and</strong> Thomas<br />

Quarries), each of which has a series of coupled <strong>marine</strong>aeolian<br />

deposits that are thought to represent individual<br />

interglacial periods. Some of the units are<br />

associated with important archaeological <strong>and</strong><br />

palaeoanthropological finds. The <strong>chronologies</strong> obtained<br />

from this study are ambiguous, with significant<br />

age reversals occurring when a MAAD OSL procedure<br />

was used <strong>and</strong> younger than expected ages were obtained<br />

using conventional, component-resolved or<br />

slow-component SAR OSL. The study, however, is<br />

interesting in regard to the application of componentresolved<br />

OSL, where one of the slow components in<br />

quartz (S3 of Singarayer & Bailey 2003) was used to<br />

extend the age range of luminescence dating. The experimental<br />

details of this dating technique are provided<br />

in Singarayer & Bailey (2003) <strong>and</strong> Singarayer (2002).<br />

The oldest age obtained was 948599 kyr, not dissimilar<br />

to a more precise, conventional SAR age of<br />

989208 kyr. The latter, however, is questionable, since<br />

the D e value is clearly obtained from a fully saturated<br />

dose response curve (see Rhodes et al. 2006: fig. 9b) <strong>and</strong><br />

this is not recommended. The slow component age is<br />

also very imprecise (60% relative error), which limits<br />

its practical use. At lower doses, its precision is comparable<br />

to that of SAR, <strong>and</strong> in many cases improved<br />

due to the linear growth of this component to much<br />

higher doses than can be achieved using the fast component<br />

<strong>and</strong> conventional SAR procedures. But its use<br />

is limited to those <strong>sediments</strong> (i.e. aeolian) that were


BOREAS <strong>Luminescence</strong> <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> 511<br />

Table 1. Comprehensive list of all luminescence dating studies applied to <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> deposits published in readily available peerreviewed<br />

journals. The appropriate reference sources are provided <strong>for</strong> each study <strong>and</strong> are matched with a specific location or region assigned a<br />

reference number that is also shown on the map of the world in Fig. 1. Also provided are the types of <strong>coastal</strong> or <strong>marine</strong> deposits dated, the<br />

luminescence method(s) used in the study or studies <strong>and</strong> the relevant time epoch. The latter is abbreviated to EP (Early Pleistocene), MP<br />

(Middle Pleistocene), LP (Late Pleistocene), H (Holocene), EH (early Holocene), MH (mid-Holocene) <strong>and</strong> LH (late Holocene). Finally, each<br />

study or group of studies was assigned to one or more of 10 broadly defined themes that best describe the overall aim(s) of the study. These<br />

themes are: A = sea-level change; B = regional-scale periods of increased aeolian activity; C = neotectonic changes; D = development of spits<br />

or barrier isl<strong>and</strong>s; E = timing <strong>and</strong> frequency of increased storminess; F = timing <strong>and</strong> recurrence intervals of tsunamis; G = reconstruction of<br />

the spatial <strong>and</strong> temporal development of local coastlines; H = effects of global warming on ‘soft’ coasts; I = archaeology <strong>and</strong> palaeoanthropology;<br />

<strong>and</strong> J = luminescence methodology studies.<br />

Ref No. Location Deposit types Age range Method Theme Reference<br />

Africa<br />

South Africa<br />

1 Langebaan aeolianite LP IRSL J Roberts & Berger (1997)<br />

2 Dias Beach dune (aeolianite) MP TL G Shaw et al. (2001)<br />

3 Die Kelders dune (cave) LP TL, OSL,<br />

IRSL<br />

J Feathers & Bush (2000)<br />

4 Agulhas Plain dunes MH-LH OSL B, G Carr et al. (2007)<br />

5 Still Bay dunes (aeolianite,<br />

cave) palaeosols<br />

LP OSL I, J Roberts et al. (2008);<br />

Henshilwood et al.<br />

(2002); Jacobs et al.<br />

(2003a, b, 2006)<br />

6 Mossel Bay dune (cave) LP OSL I Marean et al. (2007)<br />

7 Agulhas aeolianites LP OSL B Carr et al. (2007)<br />

8 Agulhas &<br />

aeolianites1barrier MP-LP OSL B, G Bateman et al. (2004)<br />

Wilderness<br />

dunes<br />

9 Wilderness barrier dunes MP-LP TL G Illenberger (1996)<br />

10 Klasies River cave dune LP TL, OSL, I Feathers (2002)<br />

IRSL<br />

11 East London aeolianites1beach<br />

rocks<br />

LP OSL A, I Jacobs & Roberts (in<br />

press)<br />

East London aeolianite LP TL, IRSL J Vogel et al. (1999)<br />

12 Maputal<strong>and</strong> aeolianite1beach MP, LP, H TL, IRSL B, G Porat & Botha (2008)<br />

rocks<br />

13 Hlabane to<br />

St.Lucia<br />

dunes LP TL, IRSL G Sudan et al. (2004)<br />

Mozambique<br />

14 Inhaca <strong>and</strong><br />

Bazaruto<br />

aeolianite1beach<br />

rock<br />

LP, H OSL A, B, D, G Armitage et al. (2006)<br />

Morocco<br />

15 Casablanca aeolian EP, MP, H OSL, TL A, I, J Rhodes et al. (2006)<br />

Australia<br />

Western Australia<br />

16 Rottnest Isl<strong>and</strong>,<br />

Perth & Point Peron<br />

17 North west Cape,<br />

Coral Bay &<br />

Shark Bay<br />

18 Cape Range Peninsula,<br />

WA<br />

19 Cape St. Lambert, East<br />

Kimberleys<br />

aeolianites LP TL B Price et al. (2001)<br />

aeolianites LP TL B Kendrick et al. (1991)<br />

<strong>marine</strong> LP, H OSL J Olley et al. (2004a)<br />

dunes MH-LH TL B Lees et al. (1992)<br />

South Australia<br />

20 Eyre Peninsula aeolianites o630 ka TL A, B Belperio (1995)<br />

21 Spencer Gulf <strong>marine</strong> <strong>sediments</strong> LP, H TL J Smith et al. (1982)<br />

22 South-east South<br />

Australia<br />

dune cordons MP-LP IRSL, TL,<br />

OSL<br />

A, B, C, G Huntley et al. (1983,<br />

1985a, b, 1993a, b,<br />

1994a, b, 1996); Huntley<br />

& Prescott (2001);<br />

Yoshida et al. (2000);<br />

Banerjee et al. (2003);<br />

Murray-Wallace et al.<br />

(1996, 1999)<br />

23 Guichen Bay beach ridges H OSL G Murray-Wallace et al.<br />

(2002a); Bristow &<br />

Pucillo (2006)


512 Zenobia Jacobs BOREAS<br />

Table 1 (continued)<br />

Ref No. Location Deposit types Age range Method Theme Reference<br />

24 Normanville embayment fill LP TL C Bourman et al. (1999)<br />

Victoria<br />

25 Point Ritchie aeolianite LP TL B, J Sherwood et al. (1994)<br />

26 Gippsl<strong>and</strong> Lakes region <strong>coastal</strong> barriers LP TL G Bryant & Price (1997)<br />

27 Nepean Peninsula aeolianites LP TL B Zhou et al. (1994)<br />

New South Wales<br />

28 Lord Howe Isl<strong>and</strong> aeolianites LP TL A, B Woodroffe et al. (1995);<br />

Price et al. (2001);<br />

Woodroffe et al. (2006)<br />

29 Southern coast tsunami LP, H TL, OSL F Price et al. (1999); Switzer<br />

et al. (2006); Young<br />

et al. (1995, 1996, 1997);<br />

Bryant et al. (1996,<br />

1997b)<br />

30 Gillards Beach & beach <strong>sediments</strong> LP TL A Young et al. (1993a)<br />

Middle Lagoon<br />

31 Steamers Beach aeolianite LP, H TL B Wheeler (1995)<br />

32 Entire coastline <strong>marine</strong> s<strong>and</strong>s MP, LP TL A Thom et al. (1994);<br />

North coast <strong>for</strong>edunes H OSL G<br />

Bryantet al. (1997a);<br />

Goodwin et al. (2006)<br />

33 Illawara coast shore plat<strong>for</strong>ms &<br />

aeolian dunes<br />

LP TL A, B Brooke et al. (1994);<br />

Bryant et al. (1990)<br />

34 Central & South coast transgressive dunes H TL G Young et al. (1993b)<br />

35 S<strong>and</strong>on Point raised <strong>marine</strong><br />

deposits<br />

LP, H TL A Bryant et al. (1992)<br />

Queensl<strong>and</strong><br />

36 Cooloola & North<br />

Stradbrooke<br />

s<strong>and</strong> dunes EP TL B Tejan-Kella et al. (1990);<br />

Yoshida et al. (2000)<br />

OSL J<br />

37 Moreton Bay indurated s<strong>and</strong>s LP, H TL, OSL G Brooke et al. (2008b)<br />

38 Keppel Bay beach1estuarine H OSL G Brooke et al. (2006;<br />

2008c); Bostock et al.<br />

(2007)<br />

Fraser Isl<strong>and</strong> beach H OSL Boyd et al. (2008)<br />

39 Northern Queensl<strong>and</strong> beach ridge1storm H OSL J Olley et al. (2004b)<br />

surges<br />

40 Ramsay Bay, North barrier dunes H TL G Pye & Rhodes (1985)<br />

Queensl<strong>and</strong><br />

41 Western Cape York <strong>marine</strong> & dune LP TL A, B Lees et al. (1993)<br />

42 Gulf of Carpentaria <strong>marine</strong> LP TL, OSL A Chivas et al. (2001);<br />

Shulmeister et al.<br />

(1993)<br />

Northern Territory<br />

43 Cape Arnhem dune podzols LP, H TL B Lees et al. (1990, 1995)<br />

44 Coburg Peninsula <strong>coastal</strong> l<strong>and</strong><strong>for</strong>ms LP TL G Woodroffe et al. (1992)<br />

Remote Oceania<br />

45 Great Barrier Isl<strong>and</strong> tsunami deposits LH IRSL F Nichol et al. (2003)<br />

46 Bream Bay beach ridge LP IRSL G Nichol (2002)<br />

47 Canterbury, Banks lake <strong>and</strong> lagoon MP-LP, H TL C, G Shulmeister et al. (1999);<br />

Peninsula<br />

48 Leithfield, North<br />

Canterbury<br />

dune ridges H TL G Shulmeister & Kirk<br />

(1996)<br />

49 Otago loess <strong>and</strong> s<strong>and</strong> LP OSL F Kennedy et al. (2007)<br />

50 Otago <strong>marine</strong><br />

terrace1alluvial<br />

fan<br />

LP IRSL, TL C Litchfield & Lian<br />

(2004); Rees-Jones<br />

et al. (2000)<br />

51 Knights Point,<br />

Westl<strong>and</strong><br />

<strong>marine</strong> LP OSL C Cooper & Kostro<br />

(2006)<br />

52 Koputaroa, North <strong>coastal</strong> dunes LP IRSL G Duller (1996)<br />

Isl<strong>and</strong><br />

53 Fiji aeolian dunes H OSL J Anderson et al. (2006)


BOREAS <strong>Luminescence</strong> <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> 513<br />

Table 1 (continued)<br />

Ref No. Location Deposit types Age range Method Theme Reference<br />

Americas<br />

South America<br />

54 Peru & Ecuador <strong>marine</strong> terraces MP-LP IRSL A, C Pedoja et al. (2006a, b)<br />

55 Southern Brazil aeolianites MP OSL, TL A, B Giannini et al. (2007)<br />

56 Brazil <strong>marine</strong> terraces LP TL, OSL A, C Baretto et al. (2002)<br />

57 Brazil <strong>marine</strong> & aeolian LP, H TL, OSL A, B Tatumi et al. (2003)<br />

United States of America <strong>and</strong> Canada<br />

58 Gulf of Mexico<br />

(Florida to Texas)<br />

<strong>coastal</strong> plain terraces<br />

& dunes<br />

MP-LP TL, OSL G Otvos (2004, 2005); Blum<br />

et al. (2002)<br />

59 Florida storm, tidal, beach modern TL J Rink & Pieper (2001)<br />

60 Florida beach ridges LH OSL D, G Lopez & Rink (2007,<br />

2008)<br />

61 Florida dunes, midden LH OSL G, I Thompson et al. (2007)<br />

62 Florida beach ridges H OSL E Rink & Forrest (2005)<br />

63 North Carolina &<br />

Virginia<br />

back-barrier dunes LH TL, IRSL,<br />

OSL<br />

B Havholm et al. (2004);<br />

Berger et al. (1991,<br />

2003)<br />

64 North Carolina beach ridges LP-H OSL B Mallinson et al. (2008)<br />

65 Massachusetts beach ridge MH IRSL A Van Heteren et al. (2000)<br />

66 Maine storm deposits LH OSL E Buynevich et al. (2007)<br />

67 New Brunswick s<strong>and</strong>y spit H IRSL D Ollerhead et al. (1994);<br />

Ollerhead & Davidson-<br />

Arnott (1995)<br />

68 Hudson Bay raised <strong>marine</strong> LP TL A Forman et al. (1987)<br />

<strong>sediments</strong><br />

69 Tuktoyaktuk, NWT, aeolian dunes LP OSL G Bateman & Murton (2006)<br />

Canada<br />

70 Alaska <strong>marine</strong> s<strong>and</strong> LP IRSL A Kaufman et al. (2001)<br />

71 Washington State tsunami LH IRSL F Huntley & Clague (1996)<br />

&British Columbia<br />

72 Oregon <strong>coastal</strong> dune s<strong>and</strong>s H OSL G Jungner et al. (2001)<br />

73 Oregon tsunami H IRSL F Ollerhead et al. (2001)<br />

74 Oregon, Cali<strong>for</strong>nia <strong>coastal</strong> dunes,<br />

estuarine<br />

LP, H TL B Peterson et al. (2007);<br />

Berger et al. (1991)<br />

Asia<br />

75 Japan, Kurosaki <strong>marine</strong> terrace LP TL, OSL C, J Tanaka et al. (1997)<br />

76 Korea, south-eastern <strong>marine</strong> terraces LP OSL, ITL C, J Choi et al. (2003a, b; 2006)<br />

peninsula<br />

77 Korea, west coast tidal flat LH IRSL A Hong et al. (2003)<br />

78 China, Bohai Bay oyster reef H OSL, IRSL A Zhang et al. (2007)<br />

79 Taiwan, south-western fluvial deposits LP, H OSL C Chen et al. (2003)<br />

<strong>coastal</strong> plain<br />

80 China, Fujia <strong>and</strong> west<br />

Guangdong<br />

‘‘old red s<strong>and</strong>s’’ H OSL B Wu et al. (2000); Zhang<br />

et al. (2008)<br />

81 Hong Kong inner shelf <strong>sediments</strong> MP-H TL, OSL, A Yim et al. (2002, 2008)<br />

IRSL<br />

82 Vietnam, south-eastern<br />

coast<br />

<strong>coastal</strong> barrier LP TL G Murray-Wallace et al.<br />

(2002b)<br />

83 Malaysia, Perak alluvium LP TL G Kamaludin et al. (1993)<br />

84 India, Bay of Bengal tsunami modern OSL F, J Murari et al. (2007)<br />

85 Sri Lanka <strong>coastal</strong> dunes LP TL B Singhvi et al. (1986)<br />

Middle East<br />

Oman<br />

86 Arabian Sea deep-sea <strong>marine</strong> LP-H OSL J Stokes et al. (2003)<br />

s<strong>and</strong>s<br />

87 Wahiba S<strong>and</strong>s <strong>coastal</strong> s<strong>and</strong>s LP, H IRSL A, B Preusser et al. (2005)<br />

United Arab Emirates<br />

88 Dubai <strong>marine</strong> H OSL G, J Z<strong>and</strong>er et al. (2007)<br />

Mediterranean<br />

Israel<br />

89 Northern <strong>coastal</strong> plain aeolianite1palaeosols<br />

LP IRSL, OSL A, B, Sivan & Porat (2004)


514 Zenobia Jacobs BOREAS<br />

Table 1 (continued)<br />

Ref No. Location Deposit types Age range Method Theme Reference<br />

90 Carmel <strong>coastal</strong> plain aeolianites1beach LP IRSL A, B, Frechen et al. (2004)<br />

rocks<br />

91 Carmel coast nearshore <strong>marine</strong> sed LP-H IRSL, TL A Porat et al. (2003)<br />

92 Central <strong>coastal</strong> plain aeolianites <strong>and</strong> LP, H IRSL B Porat et al. (2004)<br />

palaeosols<br />

93 Sharon <strong>coastal</strong> plain aeolianites LP, H IRSL B Frechen et al. (2002)<br />

94 Netanya aeolianite1palaeosols<br />

LP, H IRSL, TL B Engelmann et al.<br />

(2001)<br />

95 Givat Olga aeolianites LP, H IRSL, TL B Frechen et al. (2001)<br />

96 Tel Aviv aeolianite LP IRSL B Porat & Wintle (1995)<br />

North Africa<br />

97 Egypt, Arab Gulf Coast aeolianites LP, H OSL B El-Asmar & Wood<br />

(2000)<br />

98 Tunisia, Hergla & <strong>marine</strong>1tsunami LP OSL A, F Wood (1994)<br />

Chebba<br />

99 Tunisia beach1aeolian LP OSL J Nathan & Mauz<br />

100 Morocco, Teutan &<br />

Kebdana<br />

¨<br />

(2008)<br />

aeolianites LP TL B Bruckner (1986)<br />

Italy<br />

101 Crotone peninsula <strong>marine</strong> terraces LP IRSL, TL C Mauz & Hassler (2000)<br />

102 Calabrian peninsula <strong>marine</strong> terraces LP TL C Balescu et al. (1997)<br />

103 Tyrrhenian coast dune ridges LH IRSL G Rendell et al. (2007)<br />

104 Tyrrhenian coast <strong>marine</strong> terraces LP, H IRSL C Mauz (1999)<br />

105 Ligurian coast beach LP OSL A, C Federici & Pappalardo<br />

(2006)<br />

106 Sicily <strong>marine</strong> terraces LP TL C Mauz et al. (1997);<br />

Antonioli et al.<br />

(2006)<br />

Spain<br />

107 Tarragona aeolianite LP TL C Bruckner (1986)<br />

108 Mallorca aeolianites MP OSL KJ Nielsen et al. (2004)<br />

109 Valencia aeolianite LP TL B Fumanal (1995)<br />

110 Huelva coast aeolian<br />

dunes1palaeosols<br />

LP OSL G Zazo et al. (2005)<br />

Portugal<br />

111 Boca do Rio, Algarve tsunami modern IRSL F Dawson et al. (1995)<br />

112 Western coast aeolian dune H IRSL E, G Clarke & Rendell<br />

(2006); De Carvalho<br />

et al. (2006);<br />

Thomas et al. (2008)<br />

Northwestern Europe<br />

France<br />

113 Sangatte raised beaches MP TL A, C Balescu et al. (1991,<br />

1992)<br />

114 Aquitaine coast aeolian dunes H IRSL E, G Clarke et al. (1999;<br />

2002)<br />

115 Northern Brittany gravel1loess LP OSL G Regnauld et al. (2003)<br />

116 Norm<strong>and</strong>y <strong>marine</strong> s<strong>and</strong>s MP-LP OSL A, C Coutard et al. (2006)<br />

Engl<strong>and</strong><br />

117 Scilly Isles tsunami historical OSL F Banerjee et al. (2001)<br />

118 Dungeness beach ridges H OSL G Roberts & Plater<br />

(2007)<br />

119 North Norfolk coast intertidal <strong>sediments</strong><br />

& s<strong>and</strong> dunes<br />

H IRSL G Boomer & Horton<br />

(2006); Knight et al.<br />

(1998)<br />

120 East Yorkshire aeolian1beach LP TL, OSL G Bateman & Catt (1996)<br />

121 Northumberl<strong>and</strong> coast <strong>coastal</strong> dunes H IRSL G Wilson et al. (2001)<br />

122 Tees estuary estuarine H TL G Plater & Poolton<br />

(1992)<br />

123 Sefton coast aeolian H OSL G Pye et al. (1995)<br />

Range of locations variety modern variety J Richardson (2001)


BOREAS <strong>Luminescence</strong> <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> 515<br />

Table 1 (continued)<br />

Ref No. Location Deposit types Age range Method Theme Reference<br />

Scotl<strong>and</strong><br />

124 Orkney Isles aeolian1storm<br />

deposits<br />

LH OSL B, E Sommerville et al.<br />

E, F<br />

F<br />

(2003, 2007); Hall<br />

et al. (2006);<br />

Hansom & Hall<br />

(2008)<br />

125 Shetl<strong>and</strong> Isles s<strong>and</strong>s LP IRSL, TL B Hall et al. (2002)<br />

Irel<strong>and</strong><br />

126 Northeast coast beach ridges1dunes H IRSL A Or<strong>for</strong>d et al. (2003)<br />

127 North coast <strong>coastal</strong> dunes H IRSL, OSL B, E Wilson et al. (2004)<br />

128 Dingle Bay dune s<strong>and</strong>s LH IRSL B Wintle et al. (1998)<br />

Wales<br />

129 Aberffraw <strong>coastal</strong> dunes LH OSL B Bailey et al. (2001)<br />

Denmark<br />

130 Skallingen Spit barrier-spit dunes LH OSL A, D, E Aagaard et al. (2007)<br />

131 Wadden Sea tidal LH OSL H, J Madsen et al.<br />

(2007a, b)<br />

132 Wadden Sea estuarine LH OSL A, H, J Madsen et al. (2005)<br />

133 Jutl<strong>and</strong> beach ridge1aeolian<br />

s<strong>and</strong><br />

H OSL A, G Nielsen et al. (2006);<br />

Clemmensen et al.<br />

(2001a, 2006)<br />

B, E Murray &<br />

Clemmensen (2001)<br />

Strickertsson et al.<br />

(2001)<br />

134 Skagerrak & Kattegat aeolian dunes LH OSL B, E Clemmensen &<br />

Murray (2006)<br />

135 Gammelmark <strong>marine</strong>1aeolian LP OSL A, J Murray & Funder (2003)<br />

136 Northwestern Jutl<strong>and</strong> s<strong>and</strong> dunes LH OSL B, E Clemmensen et al. (2001b)<br />

Germany<br />

137 North Sea tidal flat LH OSL A, H Mauz & Bungenstock<br />

(2007)<br />

Netherl<strong>and</strong>s<br />

138 Texel Isl<strong>and</strong> <strong>coastal</strong> dunes modern OSL G, H, J Ballarini et al. (2003,<br />

2007); Van Heteren<br />

et al. (2006)<br />

G<br />

Lithuania<br />

139 Baltic Sea <strong>marine</strong> <strong>sediments</strong> H IRSL G Bitinas et al. (2001)<br />

Oceans <strong>and</strong> Seas<br />

140 Baltic Sea <strong>marine</strong> H OSL A Kortekaas et al. (2007)<br />

Arctic Sea deep-sea <strong>sediments</strong> LP-H OSL J Jakobsson et al. (2003)<br />

Arctic Sea <strong>marine</strong> LP-H IRSL J Berger (2006)<br />

Arctic Sea continental shelf H TL J Berger et al. (1984)<br />

North Pacific &<br />

Antarctic Oceans<br />

deep-sea <strong>sediments</strong> LP TL J Wintle & Huntley<br />

(1979a, b, 1980)<br />

18 Northwest Australia deep-sea <strong>sediments</strong> LP-H OSL J Olley et al. (2004a)<br />

86 Arabian Sea deep-sea <strong>sediments</strong> LP-H OSL J Stokes et al. (2003)<br />

exposed <strong>for</strong> a long duration (i.e. several days) to sunlight<br />

only (Singarayer et al. 2000).<br />

At the other end of Africa, along the southeastern<br />

African coast, Porat & Botha (2008) reported 57 ages<br />

from which they constructed a luminescence chronology<br />

<strong>for</strong> dune <strong>for</strong>mation on the Maputal<strong>and</strong> <strong>coastal</strong><br />

plain since <strong>marine</strong> isotope stage (MIS) 11. They used a<br />

MAAD TL procedure applied to potassium-rich feldspar<br />

grains to date beyond the effective limits of quartz<br />

OSL dating, <strong>and</strong> used a MAAD or SAAD (singlealiquot<br />

additive dose) IRSL method <strong>for</strong> the younger<br />

samples (o100 kyr). The oldest finite MAAD TL age<br />

obtained was 20313 kyr, similar to the oldest MAAD<br />

IRSL age of 21622 kyr. The advantage of TL over<br />

IRSL to extend the age range <strong>for</strong> potassium-rich feldspars<br />

was there<strong>for</strong>e not realized.<br />

Porat & Botha (2008) obtained ages <strong>for</strong> the Isipingo<br />

Formation, among others, which represents the


516 Zenobia Jacobs BOREAS<br />

25° 155° 85°<br />

Greenl<strong>and</strong><br />

Icel<strong>and</strong><br />

Asia<br />

North America<br />

Europe<br />

Japan<br />

23.5°<br />

Africa<br />

South America<br />

0°<br />

Australia<br />

23.5°<br />

New Zeal<strong>and</strong><br />

25° 155°<br />

85°<br />

Fig. 1. Map of the world indicating all study locations where luminescence dating has been used to obtain <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> or <strong>marine</strong><br />

deposits, referred to in Table 1. Green squares, red stars <strong>and</strong> blue circles designate sites of TL, IRSL <strong>and</strong> OSL dating studies, respectively.<br />

aeolianites limited to the <strong>coastal</strong> barrier. They <strong>and</strong> Armitage<br />

et al. (2006) have shown that some, but not all, of<br />

the <strong>coastal</strong> aeolianite outcrops are associated with the<br />

MIS 5 sea-level maxima <strong>and</strong> interstadials. Porat &<br />

Botha’s age estimate of 20313 kyr is consistent with an<br />

earlier, MIS 7, sea-level highst<strong>and</strong>. Perhaps more<br />

surprising is the age estimate of 17924 kyr <strong>and</strong> a number<br />

of estimates around 65 kyr, which are associated with<br />

the transitions from MIS 7 to 6 <strong>and</strong> MIS 5 to 4, respectively.<br />

At both stages, the sea level was between 60 <strong>and</strong><br />

80 m lower than at present (e.g. Waelbroeck et al. 2002).<br />

The general pattern of ages, clustered around the last<br />

few interglacial sea-level maxima, is similar to what has<br />

been observed along the southern Cape coast of South<br />

Africa. A number of studies here have used SAR OSL<br />

methods to determine the timing of aeolianite <strong>and</strong><br />

beachrock <strong>for</strong>mation (e.g. Jacobs et al. 2003a, b, 2006;<br />

Bateman et al. 2004; Jacobs & Roberts in press) to assess<br />

the association between dune <strong>for</strong>mation <strong>and</strong> changes<br />

in sea level. OSL ages obtained <strong>for</strong> beachrock at four<br />

locations along the coast at Nahoon in the Eastern<br />

Cape (Jacobs & Roberts in press), <strong>and</strong> at Sedgefield,<br />

Great Brak River <strong>and</strong> Agulhas, resulted in a combined<br />

weighted mean age estimate of 1283 kyr. Beachrock<br />

provides direct evidence of a <strong>for</strong>mer sea-level highst<strong>and</strong>,<br />

<strong>and</strong> these age estimates are consistent with independent<br />

estimates of 1261.7 kyr (based on a compilation of<br />

U-series ages) <strong>for</strong> the MIS 5e maximum (Waelbroeck<br />

et al. 2008). At these sites, aeolianites con<strong>for</strong>mably<br />

overlie the beachrock <strong>and</strong> ages obtained <strong>for</strong> these aeolianites<br />

range between 115 <strong>and</strong> 125 kyr, which confirms<br />

their <strong>marine</strong>-aeolian association. In some sedimentary<br />

sequences, there are also additional aeolianite members<br />

with ages clustering around 90–100 kyr <strong>and</strong> 80 kyr.<br />

These are, in most cases, separated from the MIS 5e<br />

aeolianite by a protosol, indicating a period of quiescence<br />

followed by dune <strong>for</strong>mation during the MIS 5c<br />

<strong>and</strong> 5a interstadials. This pattern is also reflected in the<br />

dating of dune s<strong>and</strong>s trapped in the Pinnacle Point Cave<br />

13B archaeological site (Marean et al. 2007). Here, two<br />

dune-<strong>for</strong>ming events have been dated using OSL to<br />

120 kyr <strong>and</strong> 90 kyr. The youngest of these is capped<br />

by a flowstone which has been dated by U-series to<br />

90 kyr, verifying the veracity of the OSL ages <strong>and</strong> the<br />

general patterns observed along the coast.<br />

In detail, however, the pattern is complex, reflecting<br />

the observations of Porat & Botha (2008) along the<br />

Maputal<strong>and</strong> coast <strong>and</strong> other studies further afield in the<br />

Mediterranean <strong>and</strong> Australia. There is some evidence<br />

of pulses of dune <strong>for</strong>mation during glacial periods. An<br />

example is the stack of aeolianite preserved along <strong>and</strong><br />

below the cliff face at Blombos Cave, an important archaeological<br />

site near the town of Still Bay in South<br />

Africa. These aeolianites were dated to the transition<br />

between MIS 5 <strong>and</strong> 4 at 70 kyr, <strong>and</strong> their age equivalent<br />

is found as an uncemented dune s<strong>and</strong> overlying<br />

Middle Stone Age (MSA) archaeological layers inside<br />

the cave (Jacobs et al. 2003a, b, 2006). This 70 kyr<br />

dune-<strong>for</strong>ming event has also been recorded in the<br />

Wilderness seaward dune cordon by Bateman et al.<br />

(2004). Further dune pulses during glacial periods are<br />

also recorded during MIS 3 at 392 kyr (Carr et al.


BOREAS <strong>Luminescence</strong> <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> 517<br />

2007) <strong>and</strong> MIS 6 at 160 kyr <strong>and</strong> 175–180 kyr. As a<br />

first pass comparison with the mean global sea-level<br />

curve of Waelbroeck et al. (2002), it is interesting to<br />

note that many of these dune pulses are associated with<br />

smaller-scale fluctuations in relative sea level. The continental<br />

shelf around the southern Cape coast is gently<br />

sloping <strong>and</strong> shallow, so any changes in sea level, wind<br />

direction <strong>and</strong> increased windiness may result in massive<br />

entrainment of large quantities of s<strong>and</strong>, resulting in<br />

dune <strong>for</strong>mation.<br />

The southern Cape coast of South Africa is also an<br />

important region archaeologically, where cave deposits<br />

dated using luminescence dating techniques, contain<br />

evidence of the presence of Homo sapiens (e.g. Feathers<br />

& Bush 2000; Feathers 2002) <strong>and</strong> their ability to behave<br />

in a symbolic fashion (e.g. Henshilwood et al. 2002,<br />

2004; Jacobs et al. 2006; Tribolo et al. 2006; Marean<br />

et al. 2007). The presence of Homo sapiens <strong>and</strong> other<br />

animals, such as elephants, is also known from footprints<br />

preserved in some of the aeolianite units (Roberts<br />

& Berger 1997; Roberts et al. 2008; Jacobs & Roberts in<br />

press). Underst<strong>and</strong>ing the interplay between dune <strong>for</strong>mation,<br />

fluctuating sea levels <strong>and</strong> human populations<br />

living in the <strong>coastal</strong> environment provides insights into<br />

when these cave sites would have been occupied <strong>and</strong><br />

why archaeological materials have been preserved. For<br />

example, at both Blombos Cave <strong>and</strong> Pinnacle Point<br />

Cave 13B, dunes sealed the caves at different times <strong>for</strong><br />

considerable durations (many tens of thous<strong>and</strong>s of<br />

years), making the sites unavailable as living spaces <strong>for</strong><br />

humans (Jacobs et al. 2003a, b; Marean et al. 2007).<br />

This, to some extent, explains the considerable occupational<br />

hiatuses that can be observed in the archaeological<br />

record <strong>and</strong>, at the same time, provides reasons<br />

<strong>for</strong> its good preservation.<br />

Australia<br />

The vast majority of luminescence studies that involve<br />

Australian <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> have used TL<br />

dating to derive the <strong>chronologies</strong>. There are 33 TL studies<br />

(see Table 1), 1 IRSL (Huntley et al. 1993b) <strong>and</strong> 10<br />

OSL studies (e.g. Huntley et al. 1985b, 1994b, 1996;<br />

Yoshida et al. 2000; Murray-Wallace et al. 2002a; Olley<br />

et al. 2004a, b; Brooke et al. 2006; Switzer et al. 2006;<br />

Bostock et al. 2007). These studies primarily address<br />

two broad themes: Pleistocene <strong>and</strong> Holocene dune <strong>for</strong>mation<br />

<strong>and</strong> their relation to glacial–interglacial cycles,<br />

<strong>and</strong> the timing <strong>and</strong> frequency of tsunami events.<br />

The geographic area that has received most attention<br />

with regard to investigation of the broader chronological<br />

history of Pleistocene <strong>and</strong> Holocene dune <strong>for</strong>mation<br />

in Australia is the <strong>coastal</strong> plain between<br />

Coorong <strong>and</strong> Mount Gambier in southeast South<br />

Australia (Fig. 2B). Dunes, in the <strong>for</strong>m of Holocene<br />

relict <strong>for</strong>edunes <strong>and</strong> Pleistocene relict barrier dunes,<br />

span almost the entire last 800 000 years or 25 <strong>marine</strong><br />

A<br />

δ 18 O ( )<br />

B<br />

–2<br />

–1<br />

0<br />

Robe<br />

Woakwine<br />

West Dairy<br />

East Dairy<br />

Reedy Creek<br />

West Avenue<br />

East Avenue<br />

Ardune<br />

Baker<br />

Peacock<br />

Woolumbool<br />

Stewart<br />

Harper<br />

West Naracoorte<br />

East Naracoorte<br />

100<br />

75<br />

50<br />

25<br />

0<br />

0 200 400 600 800 1000<br />

Age (kyr)<br />

WA<br />

NT<br />

Coorong<br />

SA<br />

QLD<br />

NSW<br />

V<br />

C<br />

<strong>Luminescence</strong> age (kyr)<br />

800<br />

600<br />

400<br />

200<br />

isotope stages (Fig. 2A). The Holocene dune sequences<br />

<strong>for</strong>med as a result of progradation under conditions of<br />

rapid sediment supply (Murray-Wallace et al. 2002a),<br />

whereas the Pleistocene sequence represents deposition<br />

during successive sea-level highst<strong>and</strong>s, with each barrier<br />

dune spatially separated as a result of steady tectonic<br />

uplift of the l<strong>and</strong> surface (Murray-Wallace et al.<br />

1996, 2002a) (Fig. 2A). Dating the Holocene relict<br />

dunes provides an opportunity to test how well the<br />

traps that give rise to the OSL <strong>and</strong> TL signals are<br />

bleached, <strong>and</strong> how well one can resolve ages towards<br />

the lower range of luminescence dating techniques. By<br />

contrast, the Pleistocene dune sequence offers the opportunity<br />

to test the upper limit of the dating technique,<br />

alongside its accuracy.<br />

TL <strong>and</strong> IRSL methods were applied to Holocene<br />

<strong>for</strong>edunes at Robe I (Huntley et al. 1993a, b, 1994a) <strong>and</strong><br />

Sea level (m)<br />

0<br />

0 200 600 1000<br />

Independent age (kyr)<br />

Fig. 2. A. A cross-section of the Coorong <strong>coastal</strong> plain from Robe on<br />

the coast, inl<strong>and</strong> to Naracoorte, showing the major dune cordons,<br />

some of which were dated using TL, IRSL <strong>and</strong> OSL. Also shown is a<br />

d 18 O time curve showing the matching between sea-level high st<strong>and</strong>s<br />

<strong>and</strong> corresponding dune cordons (modified from Huntley & Prescott<br />

2001). B. An outline map of Australia showing the location of the<br />

Coorong <strong>coastal</strong> plain in southeast South Australia. C. A plot of TL<br />

ages versus independent ages.


518 Zenobia Jacobs BOREAS<br />

the SAR OSL method to a series of relict <strong>for</strong>edunes on<br />

the Coorong <strong>coastal</strong> plain at Guichen Bay (Murray-<br />

Wallace et al. 2002a). Effectively, modern ages were<br />

obtained <strong>for</strong> the Robe I dune using both TL <strong>and</strong> IRSL,<br />

while an age of 515 yr was measured <strong>for</strong> the modern<br />

<strong>for</strong>edune at Long Beach using OSL. Sunlight bleaching<br />

during aeolian transport there<strong>for</strong>e seems sufficient in<br />

this region. To test the ability to precisely resolve the<br />

age of young dunes, OSL dating was applied to a series<br />

of the Guichen Bay relict <strong>for</strong>edunes. Ages were largely<br />

in sequential order <strong>and</strong> demonstrated a very fast initial<br />

rate of progradation (between 5 kyr <strong>and</strong> 4 kyr), followed<br />

by a more gradual <strong>and</strong> linear rate (after 4 kyr).<br />

This study, there<strong>for</strong>e, indicated the potential of OSL<br />

dating to provide accurate <strong>and</strong> precise <strong>chronologies</strong>,<br />

resulting in useful insights into the rate <strong>and</strong> nature of<br />

<strong>for</strong>edune development.<br />

Geological in<strong>for</strong>mation linking the Pleistocene dunes<br />

to successive high sea-level st<strong>and</strong>s, <strong>and</strong> chronological<br />

in<strong>for</strong>mation in the <strong>for</strong>m of some U-series ages (Schwebel<br />

1984), amino acid racemization ages (Murray-<br />

Wallace et al. 2001) <strong>and</strong> a palaeomagnetic reversal<br />

(Brunhes-Matuyama) found between the West <strong>and</strong><br />

East Naracoorte ranges (Idnurm & Cook 1980), provided<br />

an excellent testing ground <strong>for</strong> the per<strong>for</strong>mance<br />

of a range of luminescence methods over a significant<br />

time range. Several luminescence studies have now<br />

used this dune sequence to explicitly test the reliability<br />

of various luminescence dating methods as applied to<br />

<strong>coastal</strong> <strong>sediments</strong>. Many of these studies resulted in<br />

important methodological advances that subsequently<br />

influenced the field <strong>for</strong> many years.<br />

Huntley et al. (1993a, 1994a, b) demonstrated satisfactory<br />

agreement between the MAAD TL ages on<br />

quartz <strong>for</strong> dunes deposited between 0 <strong>and</strong> 500 kyr <strong>and</strong><br />

the independent ages <strong>and</strong> d 18 O record, but uncertainties<br />

associated with both D e <strong>and</strong> dose rate <strong>for</strong> the samples<br />

from the West <strong>and</strong> East Naracoorte dunes (the oldest in<br />

the sequence) produced indefinite results (Fig. 2C).<br />

Huntley & Prescott (2001) later applied an improved<br />

methodology involving a 33 h preheat at 1601C to<br />

empty the traps that give rise to the 2801C TL peak.<br />

This adjustment to their measurement procedures resulted<br />

in superior preheat plateaus <strong>and</strong>, thus, improved<br />

reliability of the TL ages, but still did not resolve those<br />

ages older than 500 kyr. This led them to suggest that<br />

the upper limit of TL dating is 250 Gy. To better estimate<br />

the duration of the last interglacial Woakwine<br />

barrier dune, Murray-Wallace et al. (1999) also applied<br />

TL dating, but to samples that were collected explicitly<br />

from sedimentary facies that represent the transgressive<br />

phase, last interglacial maximum <strong>and</strong> the regressive<br />

phases, as well as an older aeolianite in the core of the<br />

barrier dune. They obtained ages consistent with previous<br />

TL estimates <strong>and</strong> with expectations <strong>for</strong> the transgressive<br />

<strong>and</strong> maximum phases, but the ages <strong>for</strong> the<br />

samples associated with the regressive phases <strong>and</strong> the<br />

dune core were problematic. This demonstrates the<br />

complications arising from different modes of <strong>for</strong>mation<br />

of these dunes, which is important when improved<br />

accuracy is required alongside greater precision.<br />

Huntley et al. (1993b) also applied a MAAD-IRSL<br />

method to potassium-rich feldspar inclusions in quartz<br />

<strong>for</strong> a number of those samples that they had measured<br />

using TL (Huntley et al. 1993a). The ages were consistent<br />

within errors, but systematically younger (by less<br />

than 10%), which they suggest may have been due to<br />

anomalous fading <strong>for</strong> which they did not correct.<br />

Banerjee et al. (2003) applied the now widely used SAR<br />

procedure to quartz grains <strong>and</strong> also found good agreement<br />

with independent estimates over the last 250 kyr.<br />

They attempted to date the much older West Naracoorte<br />

dune, but age underestimates <strong>and</strong> significantly different<br />

results depending on the methodology that was used<br />

were obtained. Finally, Yoshida et al. (2000) used the<br />

SAR OSL method <strong>and</strong> applied it to single grains from<br />

the Baker (460–490 kyr), West Naracoorte (800–870 kyr)<br />

<strong>and</strong> East Naracoorte (780–990 kyr) dunes. They identified<br />

a small number of ‘supergrains’, which are quartz<br />

grains that show unusually high saturation doses, <strong>and</strong><br />

obtained ages that were in agreement with independent<br />

ages <strong>for</strong> the Baker <strong>and</strong> West Naracoorte samples,<br />

but that underestimated the expected age of the East<br />

Naracoorte sample. A problem associated with these<br />

grains is that it will be difficult to discern, in the absence<br />

of independent age control, whether or not the D e values<br />

obtained are reliable, given that so few grains yielded<br />

realistic age estimates.<br />

In summary, there<strong>for</strong>e, the luminescence <strong>chronologies</strong><br />

obtained using various different methods applied to the<br />

southeast South Australian relict barrier dunes confirm a<br />

dune sequence increasing in age inl<strong>and</strong> from the coast,<br />

controlled by changes in Quaternary sea levels.<br />

Extensive tracts of aeolianite <strong>and</strong> barrier dunes are<br />

also found elsewhere along the Australian coast. TL<br />

dating applied to these deposits in locations such as the<br />

continental coast near Perth, Western Australia (Price<br />

et al. 2001), Point Richie in Victoria (Sherwood et al.<br />

1994), the Nepean Peninsula, Victoria (Zhou et al.<br />

1994) <strong>and</strong> Lord Howe Isl<strong>and</strong> (Woodroffe et al. 1995;<br />

Price et al. 2001) also suggests that aeolianite <strong>for</strong>mation<br />

occurred at the end of, or slightly after, the last interglacial,<br />

thereby linking aeolian deposition to changes in<br />

sea level. In other locations, the TL <strong>chronologies</strong> suggest<br />

more complex depositional patterns. Along the<br />

northwestern coast (Kendrick et al. 1991), Rottnest Isl<strong>and</strong><br />

(Price et al. 2001) <strong>and</strong> most of the east coast (e.g.<br />

Bryant et al. 1990, 1997a; Bryant & Price 1997), aeolian<br />

deposition has also been observed during the last glacial<br />

period (MIS 2 <strong>and</strong> 3). Dune <strong>for</strong>mation during the<br />

Pleistocene along the entire Australian coast was thus<br />

influenced by changes in sea level <strong>and</strong> provides valuable<br />

in<strong>for</strong>mation about palaeoenvironments during glacial<br />

periods when sea levels were lower.


BOREAS <strong>Luminescence</strong> <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> 519<br />

Remote Oceania<br />

Apart from a single study in Fiji (Anderson et al. 2006),<br />

all other studies in which luminescence dating procedures<br />

have been used to obtain <strong>coastal</strong> <strong>chronologies</strong><br />

relate to <strong>sediments</strong> along the coast of New Zeal<strong>and</strong>.<br />

Perhaps it is not surprising that the greatest proportion<br />

of these studies involve some aspect of tectonics. New<br />

Zeal<strong>and</strong> is located at the boundary of the Australian<br />

<strong>and</strong> Pacific Plates, resulting in significant uplift, fault<br />

activity <strong>and</strong> the occurrence of tsunamis. Flights of<br />

<strong>marine</strong> terraces or their remnants are widely distributed<br />

<strong>and</strong> have been the subject of many studies. Many of<br />

these terraces occur at elevations substantially higher<br />

than expected based on eustatic sea-level change during<br />

the late Quaternary (5 m). This suggests that such<br />

<strong>marine</strong> terraces are tectonically or isostatically uplifted.<br />

One of the key parameters needed to determine tectonic<br />

uplift rates is a reliable estimate of the age of a <strong>marine</strong><br />

terrace, <strong>and</strong> as yet there are only a few numerical age<br />

estimates <strong>for</strong> <strong>marine</strong> terraces. Cooper & Kostro (2006)<br />

obtained a last interglacial (1237 kyr) OSL age <strong>for</strong> an<br />

uplifted (113 m a.m.s.l.) <strong>marine</strong> shoreline deposit at<br />

Knights Point, Westl<strong>and</strong>, resulting in a time-integrated<br />

uplift rate of 0.86 mm/yr relative to the Australian<br />

plate.<br />

On the east coast of South Isl<strong>and</strong>, in <strong>coastal</strong> Otago,<br />

Litchfield & Lian (2004) collected samples from raised<br />

beach s<strong>and</strong>s in close proximity to the Akatore Fault<br />

system. They obtained ages of 11712 kyr <strong>and</strong><br />

11713 kyr using IRSL <strong>and</strong> TL dating of quartz, respectively,<br />

suggesting terrace <strong>for</strong>mation during MIS 5e<br />

(sea-level maximum) at a location only 300 m from the<br />

Akatore Fault. The terrace is only 4 m a.m.s.l., suggesting<br />

negligible uplift since that time. These ages are<br />

different from the age estimate of 7114 kyr obtained<br />

<strong>for</strong> the same <strong>marine</strong> s<strong>and</strong>s by Rees-Jones et al. (2000),<br />

but the discrepancy is thought to be due to the use of an<br />

inappropriate procedure, i.e. optical dating of feldspar<br />

inclusions in quartz grains (e.g. Huntley et al. 1993b) by<br />

the latter authors. Further away (10 km) from the influence<br />

of faulting, IRSL ages of 9711 kyr <strong>and</strong><br />

965 kyr were calculated <strong>for</strong> raised beach s<strong>and</strong>s<br />

(Litchfield & Lian 2004). These ages were not corrected<br />

<strong>for</strong> anomalous fading either, so it is difficult to verify<br />

whether this <strong>marine</strong> terrace is associated with MIS 5e or<br />

one of the other MIS 5 interstadials <strong>and</strong>, hence, the<br />

implications <strong>for</strong> estimation of uplift rates. Nonetheless,<br />

the 117 kyr ages <strong>for</strong> the beach s<strong>and</strong> closest to the Akatore<br />

fault suggest that no significant tectonic activity<br />

occurred on the Akatore fault since the last interglacial<br />

125 kyr ago. Similar tectonic stability on the east<br />

coast of South Isl<strong>and</strong> has been observed by Shulmeister<br />

et al. (1999), who took samples <strong>for</strong> TL dating from<br />

nearly the entire length of a 75 m core collected from<br />

the Banks Peninsula, near Christchurch. The core contains<br />

a 200 kyr sequence of alternating mud, loess <strong>and</strong><br />

soil lenses. TL ages suggest the mud relates to the last<br />

three interglacial periods <strong>and</strong> the soil to the intervening<br />

stadials. Based on existing sea-level curves <strong>for</strong> MIS 5, it<br />

has been shown by this study that the Banks Peninsula<br />

has been tectonically stable <strong>for</strong> at least the last 125 kyr.<br />

Tsunami <strong>and</strong> storm-surges are relatively common in<br />

New Zeal<strong>and</strong>. Two different types of tsunami deposits<br />

have been dated using OSL from polymineral fine<br />

grains (Nichol et al. 2003; Kennedy et al. 2007). Nichol<br />

et al. (2003) obtained OSL ages on dune s<strong>and</strong>s underlying<br />

<strong>and</strong> overlying a gravel sheet extending beyond<br />

(14.3 m) the limits of storm surge (0.8 m) on Great<br />

Barrier Isl<strong>and</strong>, located off the west coast of North Isl<strong>and</strong>.<br />

The ages bracket the tsunami to between 0 <strong>and</strong><br />

4.7 kyr, but are unable to provide better resolution.<br />

Along the north Otago coast, South Isl<strong>and</strong>, Kennedy<br />

et al. (2007) described deposits of large, imbricated<br />

boulders elevated above modern sea level. They dated<br />

s<strong>and</strong> grains found in between the boulders <strong>and</strong> also two<br />

units of in situ loess deposits overlying the boulders.<br />

The s<strong>and</strong> grains between the boulders <strong>and</strong> the loess directly<br />

overlying these boulders date to MIS 5a, providing<br />

a depositional age <strong>for</strong> the tsunami, the oldest known<br />

in New Zeal<strong>and</strong>.<br />

The Americas<br />

A large proportion of the studies conducted along the<br />

North American coastline concern the age of the <strong>for</strong>mation<br />

of Holocene barrier dunes. These l<strong>and</strong><strong>for</strong>ms are<br />

of interest because their stratigraphic records often<br />

show identifiable marker horizons <strong>and</strong> erosional<br />

boundaries that indicate connections between <strong>coastal</strong><br />

processes <strong>and</strong> barrier behaviour. Dating of such stratigraphic<br />

features can, <strong>for</strong> example, allow determination<br />

of the timing <strong>and</strong> recurrence intervals of storms (e.g.<br />

Rink & Forrest 2005; Buynevich et al. 2007). Using<br />

OSL dating of quartz grains, Rink & Forrest (2005)<br />

determined the mean recurrence interval of major<br />

storms along the Canaveral Peninsula in Florida to be<br />

808 years over the last 4000 years, using the average<br />

ridge accumulation rate derived from the dating of a<br />

limited number of dunes as a proxy <strong>for</strong> storm activity.<br />

Buynevich et al. (2007) obtained a more direct estimate<br />

of storm frequency along the coast of Maine during the<br />

last 1500 years. They used diagnostic geophysical <strong>and</strong><br />

sedimentological signatures of severe erosional events<br />

preserved in the barrier dune to estimate a storm recurrence<br />

interval of 100 years over the last 500 years,<br />

preceded by a calmer period lasting 1000 years.<br />

Dating of barrier dunes can also lead to the identification<br />

of regionally extensive periods of increased aeolian<br />

activity (e.g. Berger et al. 2003; Havholm et al.<br />

2004; Mallinson et al. 2008). Along the North Carolina<br />

<strong>and</strong> Virginia coasts, alternating patterns of dune activity<br />

<strong>and</strong> stabilization were observed by Havholm et al.


520 Zenobia Jacobs BOREAS<br />

(2004) using ground penetrating radar (GPR), <strong>and</strong> the<br />

periods of dune stability were dated using OSL by Berger<br />

et al. (2003). They applied a combination of quartz<br />

OSL <strong>and</strong> IRSL on feldspars, but found that the IRSL<br />

ages were too young due to the occurrence of anomalous<br />

fading; they there<strong>for</strong>e based their ages <strong>for</strong> dune<br />

<strong>for</strong>mation on the OSL results. The resultant <strong>chronologies</strong><br />

<strong>and</strong> GPR observations led them to identify<br />

three phases of dune activity, punctuated by two phases<br />

of dune stability, during the last 1800 years. Havholm<br />

et al. (2004) suggested that the conditions resulting in<br />

dune <strong>for</strong>mation <strong>and</strong> preservation in a mostly humid<br />

region were likely the result of a combination of climatic<br />

deterioration, in the <strong>for</strong>m of increased storminess<br />

<strong>and</strong> cooler temperatures, <strong>and</strong> a gradually rising sea level.<br />

A similar scenario was proposed by Jungner et al.<br />

(2001), who used OSL dating to investigate the cycles of<br />

stabilization <strong>and</strong> reactivation that <strong>for</strong>med the dune sequence<br />

observed at Cape Kiw<strong>and</strong>a, Oregon, during the<br />

last 7000 years. In other areas of North America, barrier<br />

dunes have been identified as palaeo-shorelines, so<br />

that dating them may provide in<strong>for</strong>mation about the<br />

response of barrier dunes to changes in sea level (e.g.<br />

Van Heteren et al. 2000; Lopez & Rink 2007, 2008;<br />

Mallinson et al. 2008). For a barrier in Massachusetts,<br />

Van Heteren et al. (2000) used a multiple-aliquot additive<br />

<strong>and</strong> regenerative dose IRSL procedure to obtain<br />

ages <strong>for</strong> samples collected from the full vertical range<br />

(8 m) of an aeolian unit above the dune-beach facies<br />

boundary to determine changes in sea level. After correcting<br />

<strong>for</strong> anomalous fading, they obtained ages that<br />

showed satisfactory agreement with 14 C ages obtained<br />

from a salt-marsh peat located behind the barrier dune.<br />

Their chronology suggested that local relative sea level<br />

rose 8 m during the past 5500 years. Lopez & Rink<br />

(2007) showed that, due to rising sea levels, emergent<br />

l<strong>and</strong> has <strong>for</strong>med on St. Vincent Isl<strong>and</strong>, one of the Apalachicola<br />

barrier isl<strong>and</strong>s along the northwest Gulf coast<br />

of Florida, since 4000 years ago. They obtained sequentially<br />

younger ages coastward, with differential<br />

accumulation rates during the late Holocene <strong>and</strong> in<br />

different parts of the isl<strong>and</strong>. The impact of changing sea<br />

level has also been detected from barrier deposits dating<br />

to the Late Pleistocene in North Carolina (e.g.<br />

Mallinson et al. 2008). They obtained OSL ages related<br />

to MIS 5a <strong>and</strong> MIS 3, from which they deduced that the<br />

relative sea level in this area at these times was at or<br />

above present sea level. The Holocene palaeo-shoreline<br />

at 4–3 kyr lay 2–3 km west (i.e. inl<strong>and</strong>) of the current<br />

shoreline. They suggested that although this may have<br />

resulted from rapid, minor sea-level oscillations during<br />

the late Holocene, it is more likely to represent the rapid<br />

flooding <strong>and</strong> l<strong>and</strong>ward translation of the shoreline.<br />

There<strong>for</strong>e the <strong>chronologies</strong> obtained from these barrier<br />

dunes can provide useful insights into climate cycles<br />

<strong>and</strong> isostatic movements in middle- <strong>and</strong> high-latitude<br />

regions since the last glaciation. Calculations of storm<br />

event frequencies can also influence <strong>coastal</strong> management,<br />

development <strong>and</strong> protection planning decisions.<br />

Pleistocene aeolian <strong>and</strong> <strong>marine</strong> successions have also<br />

been dated. Of particular interest are those along the<br />

South American coastline. Three studies used luminescence<br />

dating to provide <strong>chronologies</strong> <strong>for</strong> aeolian <strong>and</strong><br />

<strong>marine</strong> successions identified along the Brazilian coast<br />

(Baretto et al. 2002; Tatumi et al. 2003; Giannini et al.<br />

2007) with a further two studies along the coasts of<br />

Ecuador <strong>and</strong> Peru (Pedoja et al. 2006a, b). The Brazilian<br />

coast is located along a passive continental margin<br />

<strong>and</strong> provides an opportunity to directly assess the extent<br />

of sea-level change in the absence of tectonic influences,<br />

whereas the coasts of Ecuador <strong>and</strong> Peru are<br />

heavily uplifted due to subduction of the adjacent<br />

Carnegie Ridge. Baretto et al. (2002) used a combination<br />

of TL <strong>and</strong> OSL dating of quartz <strong>and</strong> IRSL dating<br />

of feldspar grains to obtain direct estimates of sea-level<br />

change from two <strong>marine</strong> terraces along the coast of the<br />

Rio Gr<strong>and</strong>e do Norte State in northeastern Brazil. The<br />

ages of 220–206 kyr <strong>and</strong> 117–110 kyr obtained were<br />

correlated with the sea-level highst<strong>and</strong>s of MIS 7c <strong>and</strong><br />

5e, respectively. The MIS 7c <strong>marine</strong> terrace ranges in<br />

altitude between 7.5 <strong>and</strong> 1.3 m a.m.s.l. <strong>and</strong> the MIS 5e<br />

<strong>marine</strong> terrace between 1 <strong>and</strong> 20 m a.m.s.l. The combination<br />

of elevation <strong>and</strong> age estimates suggests relative<br />

down-faulting of the MIS 7c deposit <strong>and</strong> uplift of the<br />

MIS 5e deposit. This study showed, there<strong>for</strong>e, that it is<br />

inappropriate to establish sea-level change from terrace<br />

elevation alone, even on passive continental margins;<br />

rather, numerical-age chronology is essential.<br />

Along the northwestern coast of South America, in<br />

Ecuador <strong>and</strong> Peru, coasts have been substantially uplifted.<br />

Marine terraces reaching elevations of 360 m<br />

a.m.s.l. were dated by Pedoja et al. (2006a, b) using a<br />

modified SAR IRSL procedure <strong>for</strong> feldspars (Lamothe<br />

& Auclair 1999), <strong>and</strong> anomalous fading corrections<br />

were made (following Huntley & Lamothe 2001).<br />

Relatively imprecise (5–25% error) age estimates were<br />

obtained <strong>and</strong> assigned to specific sea-level highst<strong>and</strong>s<br />

during MIS 9, 7 <strong>and</strong> 5e. Although association of <strong>marine</strong><br />

terraces to a specific highst<strong>and</strong> is not fully supported<br />

by the IRSL ages (e.g. mean ages of some of the samples<br />

are consistent with MIS 4, but ascribed to MIS 5e),<br />

relative time-integrated uplift rates of between 0.10<br />

<strong>and</strong> 0.50 mm/year were calculated (Pedoja et al.<br />

2006a, b).<br />

Asia<br />

Many parts of Asia, especially the eastern parts, are<br />

subjected to tectonic <strong>for</strong>ces today <strong>and</strong> were also tectonically<br />

very active during much of the late Quaternary.<br />

Geomorphic features such as <strong>marine</strong> terraces can be<br />

dated to determine the rate of uplift (e.g. Tanaka et al.<br />

1997; Choi et al. 2003a, b, 2006), from which tectonic


BOREAS <strong>Luminescence</strong> <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> 521<br />

stability can be deduced. Alternatively, thick basin-fill<br />

<strong>sediments</strong> can be dated to determine subsidence rates,<br />

as was done <strong>for</strong> the southwestern <strong>coastal</strong> plain of<br />

Taiwan (Chen et al. 2003). A further outcome of tectonic<br />

(seismic) activity is the generation of tsunamis,<br />

which deposit large sheets of s<strong>and</strong> or gravel in the<br />

onshore run-up zone. Several studies in Asia have<br />

addressed issues related to tectonics, <strong>and</strong> some of these<br />

have reported interesting observations of luminescence<br />

phenomena that have helped improve the accuracy <strong>and</strong><br />

precision of their age estimates. Such progress is necessary<br />

to estimate accurate uplift <strong>and</strong> subsidence rates, as<br />

well as recurrence intervals of events such as tsunamis.<br />

The timing of <strong>marine</strong> terrace <strong>for</strong>mation along the<br />

southeastern coastline of the Korean Peninsula (Choi<br />

et al. 2003a, b, 2006) <strong>and</strong> the west coast of central Japan<br />

(Tanaka et al. 1997) has been determined using luminescence<br />

dating methods. In both areas, bleaching experiments<br />

were per<strong>for</strong>med, using either modern<br />

analogues (Choi et al. 2003b, 2006) or laboratorycontrolled<br />

illumination conditions (Tanaka et al. 1997)<br />

to determine the residual signal <strong>for</strong> TL dating to confirm<br />

the well-bleached nature of <strong>sediments</strong> <strong>for</strong> optical<br />

dating. These studies determined that both TL <strong>and</strong><br />

OSL can be employed, but that the light-insensitive TL<br />

signal should be determined <strong>and</strong> the ages corrected accordingly.<br />

Tanaka et al. (1997) demonstrated satisfactory<br />

agreement of TL <strong>and</strong> OSL ages with independent<br />

ages <strong>for</strong> <strong>marine</strong> terraces in Japan, but these estimates<br />

had very large uncertainties.<br />

Choi et al. (2003b) used the SAR OSL procedure to<br />

determine ages <strong>for</strong> <strong>marine</strong> terraces in Korea. Previous<br />

OSL dating attempts had revealed problematic behaviour,<br />

resulting in age inversions <strong>and</strong> a lack of reproducibility<br />

of ages in a specific unit. Choi et al.<br />

(2003b) experienced similar problems when using a<br />

1601C cutheat <strong>for</strong> the test doses <strong>and</strong> comparatively low<br />

preheat temperatures <strong>for</strong> the natural <strong>and</strong> regenerative<br />

doses. Using linearly modulated (LM) OSL measurements,<br />

they were able to demonstrate the presence of a<br />

thermally unstable ‘ultrafast’ component (Jain et al.<br />

2003, 2008) in some of their samples, <strong>and</strong> suggested that<br />

increasing the cutheat <strong>and</strong> preheat temperatures to at<br />

least 2201C eliminated this undesirable component.<br />

The ultrafast component is not present in the natural<br />

signal, <strong>and</strong> it sensitizes at a different rate to the other<br />

components, which results in an inadequate sensitivity<br />

correction, such that preheat plateaus cannot be obtained<br />

<strong>and</strong> laboratory doses cannot be recovered in<br />

dose recovery experiments. By adapting their measurement<br />

procedures, Choi et al. (2003b) were able to obtain<br />

reproducible results <strong>for</strong> the majority of their<br />

samples, <strong>and</strong> identified evidence of two palaeo-shorelines<br />

at 50–70 kyr <strong>and</strong> 110–120 kyr <strong>for</strong> laterally discontinuous<br />

terraces extending over several tens of<br />

kilometres. There were, however, some samples in<br />

which problems persisted, but they were able to<br />

overcome these by deconvoluting the decay curves <strong>and</strong><br />

estimating the D e from the separated fast-component.<br />

This approach produced ages in correct stratigraphic<br />

order <strong>and</strong> in agreement with ages from other sections<br />

(Choi et al. 2003a).<br />

Choi et al. (2003b) did not succeed in obtaining ages<br />

<strong>for</strong> any of their poorly sorted <strong>sediments</strong>. They suggested<br />

that these may contain clasts of weathered gravel<br />

that, during in situ decomposition <strong>and</strong> laboratory<br />

h<strong>and</strong>ling, may liberate unbleached grains into the<br />

otherwise fully bleached <strong>sediments</strong>. Using single, multigrain<br />

aliquots of quartz, would thus result in an overestimation<br />

of the burial age. Choi et al. (2003b) did not<br />

attempt to make any single-grain analyses, which may<br />

have been able to identify the unbleached grains from<br />

the D e distribution. This may have allowed determination<br />

of the age from the most fully bleached grains,<br />

using statistical models such as the minimum age model<br />

of Galbraith et al. (1999).<br />

Using some of the same samples, Choi et al. (2006)<br />

also tried dating with an alternative signal, the 3101C<br />

isothermal TL signal, to determine their depositional<br />

ages. Their aim was to extend the age range, since many<br />

of the samples from these terraces are very close to saturation,<br />

owing to the relatively high dose rates<br />

(2–4 Gy/kyr). They demonstrated good agreement of<br />

ITL ages with OSL ages <strong>for</strong> some, but not all, of the<br />

<strong>marine</strong> terrace <strong>sediments</strong> <strong>and</strong> showed that the isothermal<br />

TL signal does not originate from a single trap<br />

type; this may have implications <strong>for</strong> adequate bleaching<br />

of the signal be<strong>for</strong>e deposition. They were unable to<br />

obtain a zero age <strong>for</strong> modern analogues, but the residual<br />

dose of 15 Gy would be negligible <strong>for</strong> samples<br />

with D e values of 4300 Gy, thus emphasizing the potential<br />

<strong>for</strong> long-range dating.<br />

Choi et al. (2003a, b, 2006) concluded that considerable<br />

tectonic activity, much greater than previously believed,<br />

had occurred in the area during the late<br />

Pleistocene. This was supported by the occurrence of<br />

the 50–70 kyr palaeo-shoreline at significantly different<br />

altitudes (7–25 m a.m.s.l.) across the study area.<br />

The isl<strong>and</strong> of Taiwan, currently a tectonically active<br />

area, is experiencing the opposite effect to Korea,<br />

namely significant subsidence, the rate of which can be<br />

deduced from the analysis of basin-fill <strong>sediments</strong>,<br />

mostly of fluvial origin. Chen et al. (2003) obtained<br />

stratigraphically coherent SAR OSL age estimates<br />

consistent with 14 C ages <strong>for</strong> <strong>sediments</strong> deposited during<br />

the late Pleistocene <strong>and</strong> Holocene. The dated <strong>sediments</strong><br />

were extracted from five 250 m long cores collected<br />

from the <strong>coastal</strong> plain basin-fill in western Taiwan. The<br />

results showed spatially variable tectonic activity over a<br />

relatively small area <strong>and</strong> deduced, from a comparison<br />

with global sea level, that the amount <strong>and</strong> pattern of<br />

sediment deposition reflects not only changes in sea level<br />

but also subsidence. Based on changes in rates of<br />

deposition between different cores, they argued that the


522 Zenobia Jacobs BOREAS<br />

location of the depo-centre was situated southwest of<br />

the study area. These observations are of relevance to<br />

engineering risk <strong>and</strong> <strong>coastal</strong> management in an area<br />

that is heavily populated.<br />

Large seismic events can trigger the <strong>for</strong>mation of<br />

tsunamis, which leave onshore deposits, such as sheets<br />

of s<strong>and</strong> or gravel, in the run-up zone. Many studies<br />

around the world have attempted to estimate the depositional<br />

age of such <strong>sediments</strong> using TL, OSL <strong>and</strong><br />

IRSL dating. Sufficient exposure of the transported <strong>sediments</strong><br />

to sunlight is a basic premise that should be<br />

met when dating tsunami deposits. The 2004 Boxing<br />

Day tsunami was associated with the second largest<br />

earthquake ever recorded, <strong>and</strong> this provided an opportunity<br />

to test the applicability of OSL dating to such<br />

<strong>sediments</strong>. Murari et al. (2007) used the conventional<br />

SAR OSL procedure as well as a fast-component SAR<br />

variant to obtain D e values from 9 samples collected<br />

from 15 cm below the surface of a 1 m deep s<strong>and</strong><br />

sheet deposited by this tsunami. They obtained ages of<br />

250 years using conventional SAR OSL <strong>and</strong> ages of<br />

o50 years using the separated fast component. This led<br />

them to suggest that some proportion of the OSL signal<br />

– the fast component – will be bleached during sediment<br />

transport by tsunamis, even though the daylight exposure<br />

time is limited. Furthermore, regardless of the<br />

energetic nature of a tsunami of this magnitude, they<br />

thought it likely to have scoured only the top 50 cm of<br />

deposit in the intertidal region, based on the fact that<br />

they obtained reproducible results from their OSL<br />

measurements. These indicated no mixing with<br />

1.6 kyr <strong>sediments</strong> (dated by 14 C) found at depths of<br />

45–88 cm. This study there<strong>for</strong>e shows that OSL dating<br />

can be used to date recent tsunami deposits with some<br />

accuracy using the fast component in quartz, but it<br />

sheds doubt on ages derived using the hard-to-bleach<br />

TL signal, which had been used previously in other regions<br />

of the world.<br />

There are some areas along the Asian coast that are<br />

tectonically stable <strong>and</strong> from which useful in<strong>for</strong>mation<br />

can be gleaned in regard to eustatic sea-level change.<br />

Cores composed of both <strong>marine</strong> <strong>and</strong> terrestrial <strong>sediments</strong><br />

were collected from the stable continental margin<br />

at two locations near Hong Kong. The <strong>marine</strong> <strong>and</strong><br />

terrestrial s<strong>and</strong>y <strong>sediments</strong> were dated using TL (Yim<br />

et al. 2002) <strong>and</strong> a combination of different OSL <strong>and</strong><br />

IRSL techniques on quartz <strong>and</strong> feldspar (Yim et al.<br />

2008). The age estimates are ambiguous, with Yim et al.<br />

(2008) reporting differences between different OSL <strong>and</strong><br />

IRSL methods. Some of these problems are related to<br />

dose saturation <strong>and</strong> others to anomalous fading, <strong>for</strong><br />

which no corrections were made. Neither study involved<br />

a detailed assessment of the dose rate estimates,<br />

although consideration is given to moisture content,<br />

which may have varied substantially in the past due to<br />

occasional subaerial exposure during periods of lowered<br />

sea level <strong>and</strong> also compaction. Furthermore, Yim<br />

et al. (2008) measured the parent concentrations of U<br />

<strong>and</strong> Th with neutron activation analysis (NAA). This<br />

approach is apt to result in inaccurate dose-rate estimates<br />

if significant disequilibria are present in the uranium<br />

decay chain (Olley et al. 1996). Such a scenario is<br />

quite likely in this offshore environment, <strong>and</strong> will be<br />

significant <strong>for</strong> the studied samples that have unusually<br />

high U <strong>and</strong> Th concentrations (4–5 ppm <strong>and</strong><br />

24–37 ppm, respectively). Sediments collected from<br />

such environments require assessment of the U-series<br />

equilibrium status using high-resolution gamma <strong>and</strong><br />

alpha spectrometry, which allow any time-dependent<br />

changes in the dose rate to be modelled (e.g. Z<strong>and</strong>er<br />

et al. 2007).<br />

Middle East<br />

This review revealed only two luminescence dating studies<br />

have been conducted along the non-Mediterranean<br />

<strong>coastal</strong> margin of the Middle East, excluding the application<br />

of OSL dating to two ocean cores off the Arabian<br />

coast (discussed in the section on oceans <strong>and</strong> seas,<br />

below). Z<strong>and</strong>er et al. (2007) applied SAR OSL dating<br />

procedures to samples collected from beachrock, beach<br />

ridge <strong>and</strong> s<strong>and</strong>-flat <strong>sediments</strong> exposed during archaeological<br />

excavations at two sites in the United Arab<br />

Emirates. They determined the dose rates using highresolution<br />

gamma spectrometry <strong>and</strong>, <strong>for</strong> comparison,<br />

inductively coupled plasma mass spectrometry. They<br />

observed significant 234 U disequilibrium in the <strong>marine</strong>derived<br />

deposits compared to its daughter 226 Ra, which<br />

they attribute to the likely uptake of uranium by shell<br />

hash <strong>and</strong> other carbonates found within the <strong>sediments</strong>.<br />

They modelled geochemically open <strong>and</strong> closed system<br />

scenarios to enable calculation of maximum <strong>and</strong> minimum<br />

ages <strong>for</strong> each sample. Satisfactory agreement was<br />

obtained with 14 C ages when using the open system<br />

model, assuming the steady (linear) post-mortem uptake<br />

of U in the shell debris. This dose-rate correction<br />

method allowed them to reconstruct the Holocene<br />

<strong>coastal</strong> evolution of Dubai in the United Arab<br />

Emirates. They provided chronological evidence <strong>for</strong> a<br />

<strong>marine</strong> highst<strong>and</strong> at 7.1 kyr, as reported previously by<br />

Goudie et al. (2000) in the northeast of the UAE <strong>and</strong> by<br />

Preusser et al. (2005); the latter workers determined a<br />

mean age <strong>for</strong> four beach s<strong>and</strong>s along the southeast<br />

Arabian Peninsula of 6.2 kyr, in accordance with the<br />

beginning of the Holocene eustatic high sea-level st<strong>and</strong>.<br />

The highst<strong>and</strong> in Dubai was followed by the development<br />

<strong>and</strong> then drowning of a mangrove swamp<br />

ecosystem, which was followed in turn by s<strong>and</strong> bar,<br />

barrier-isl<strong>and</strong> <strong>for</strong>mation <strong>and</strong> filling-in processes that<br />

resulted in an extension of the <strong>coastal</strong> plain. A brief<br />

<strong>marine</strong> incursion 4.2 kyr ago created an erosion<br />

plat<strong>for</strong>m that was subsequently covered by s<strong>and</strong>-flat<br />

<strong>sediments</strong> <strong>and</strong> archaeological materials.


BOREAS <strong>Luminescence</strong> <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> 523<br />

Preusser et al. (2005) dated 42 samples from the<br />

<strong>coastal</strong> deposits of the Wahiba S<strong>and</strong> Sea in Oman,<br />

using a multiple-aliquot additive-dose IRSL method, to<br />

derive palaeoenvironmental in<strong>for</strong>mation <strong>for</strong> the late<br />

Pleistocene <strong>and</strong> Holocene. In addition to the beach<br />

s<strong>and</strong>s discussed above, they collected samples from<br />

aeolianites <strong>and</strong> dune s<strong>and</strong>s. The timing of three pulses<br />

of dune <strong>for</strong>mation was identified: the transition from<br />

MIS 3 to 2 (34–24 kyr), the high-latitude late-glacial<br />

periods (16–13 kyr) <strong>and</strong> the early Holocene<br />

(9–8 kyr). Each of these represents an arid period<br />

during which s<strong>and</strong> supply increased due to exposure of<br />

the continental shelf. The aeolian dune <strong>for</strong>mations<br />

around the coast were shown not to be synchronous<br />

with those further north <strong>and</strong> inl<strong>and</strong>. This asynchroneity,<br />

<strong>and</strong> the lack of Last Glacial Maximum deposits,<br />

was interpreted by Preusser et al. (2005) as the result of<br />

deficient accumulation potential rather than reduced<br />

aeolian transport.<br />

Both the Z<strong>and</strong>er et al. (2007) <strong>and</strong> Preusser et al.<br />

(2005) studies demonstrate the potential <strong>for</strong> dating<br />

<strong>coastal</strong> deposits in the Middle East to better our underst<strong>and</strong>ing<br />

of Quaternary climates in a region that is<br />

highly sensitive to changes in the Southwest Asian<br />

monsoon system.<br />

Mediterranean<br />

A number of studies have applied luminescence dating<br />

procedures to different types of <strong>coastal</strong> deposits in the<br />

Mediterranean over the years. Most of these studies<br />

were conducted along the coasts of Italy <strong>and</strong> Israel,<br />

with a limited number in Spain, Portugal <strong>and</strong> along the<br />

Mediterranean coast of North Africa. The studies conducted<br />

in Italy were mostly concerned with the dating<br />

of uplifted <strong>marine</strong> shorelines to help reconstruct the<br />

Pleistocene sea-level evolution of the Mediterranean<br />

Basin, assess local <strong>and</strong> regional de<strong>for</strong>mation rates <strong>and</strong><br />

evaluate regional differences in the intensity of tectonic<br />

changes. Uplifted <strong>marine</strong> deposits are found along the<br />

Italian coast at elevations ranging substantially in space<br />

<strong>and</strong> time. Balescu et al. (1997) were the first to apply<br />

luminescence dating to <strong>coastal</strong> deposits in Italy, along<br />

the Calabrian coast of southern Italy. They used alkali<br />

feldspar grains to obtain TL ages <strong>for</strong> last interglacial,<br />

shallow <strong>marine</strong> deposits at two MIS 5e (Tyrrhenian)<br />

reference sites, Ravagnese <strong>and</strong> Bovetto, where deposits<br />

occur at 129 m <strong>and</strong> 125 m a.m.s.l., respectively. The TL<br />

ages of 11613 kyr <strong>and</strong> 11612 kyr obtained after correcting<br />

<strong>for</strong> anomalous fading, confirmed their MIS 5e<br />

association. Average uplift rates of 1 mm/year were<br />

calculated <strong>for</strong> these deposits. In contrast, Federici &<br />

Pappalardo (2006) reported a single OSL age <strong>for</strong> a last<br />

interglacial beach deposit at 28 m a.m.s.l. along the Ligurian<br />

coast, resulting in an order-of-magnitude smaller,<br />

<strong>and</strong> very gradual, uplift rate of 0.18 mm/year.<br />

Along the northwest coast of Sicily, Mauz et al. (1997)<br />

reported TL ages consistent with MIS 5e at elevations<br />

of between 15 <strong>and</strong> 118 m a.m.s.l. Antonioli et al.<br />

(2006) indicated that this part of Sicily is relatively<br />

stable, with little evidence of uplift during the Holocene<br />

<strong>and</strong> uplift rates of 0.08–0.56 mm/year during the late<br />

Quaternary. These variable uplift rates along the Italian<br />

coast determined on the basis of TL ages (<strong>and</strong> independent<br />

age control) indicate, there<strong>for</strong>e, that late<br />

Quaternary tectonic movements in this region have<br />

been significantly different over small distances, resulting<br />

in a very complex tectonic history.<br />

Two further detailed studies of sedimentary successions<br />

along the central west (Tyrrhenian) <strong>and</strong> southern<br />

(Crotone) Italian coasts provided additional in<strong>for</strong>mation<br />

about eustatic sea-level change spanning the period<br />

between the last interglacial <strong>and</strong> the end of the last glacial<br />

(Mauz 1999; Mauz & Hassler 2000). Both studies<br />

used MAAD TL <strong>and</strong> IRSL procedures to obtain ages<br />

<strong>for</strong> the different sedimentary units. Although the calculated<br />

ages were relatively imprecise due to anomalous<br />

fading <strong>and</strong> substantial measurement uncertainties, both<br />

studies indicated that the extent of eustatic sea-level<br />

change was smaller in the Mediterranean than in the<br />

open oceans, especially during MIS 3, highlighting the<br />

lack of in<strong>for</strong>mation on Mediterranean thermohaline<br />

circulation <strong>and</strong> sea-water exchange between the Atlantic<br />

<strong>and</strong> Mediterranean sea during glacial periods. Improved<br />

precision in age control, through the use of SAR OSL<br />

applied to these extensive deposits found along the Italian<br />

coast, could make a substantial contribution towards<br />

improving our underst<strong>and</strong>ing of sea-level change<br />

in the Mediterranean. The complex interplay between<br />

sea-level change, tectonic activity <strong>and</strong> changing environments<br />

along the Atlantic coast near Huelva, southwestern<br />

Spain, has also been documented using OSL<br />

dating (e.g. Zazo et al. 2005).<br />

Further east, on the coast of Israel, three to five narrow<br />

linear ridges running parallel to the str<strong>and</strong>line <strong>for</strong>m<br />

the most conspicuous geomorphic features on the<br />

<strong>coastal</strong> plain. These ridges consist of a series of aeolianites<br />

(kurkar), palaeosols (hamra) <strong>and</strong> beachrock,<br />

which are believed to indicate changing <strong>coastal</strong>, continental<br />

<strong>and</strong> <strong>marine</strong> environmental influences. <strong>Luminescence</strong><br />

studies have attempted to detail the<br />

stratigraphic relations between the kurkar, hamra <strong>and</strong><br />

beachrock, since this <strong>coastal</strong> sequence is an important<br />

archive of climate change <strong>and</strong> environment <strong>for</strong> this region.<br />

A large number of IRSL <strong>and</strong> TL ages have been<br />

determined <strong>for</strong> outcrops along the Carmel <strong>and</strong> Sharon<br />

<strong>coastal</strong> plains to estimate the <strong>for</strong>mation times of kurkar<br />

<strong>and</strong> hamra (e.g. Engelmann et al. 2001; Frechen et al.<br />

2001, 2002, 2004; Porat et al. 2004). MAAD IRSL <strong>and</strong><br />

TL ages were calculated <strong>for</strong> outcrops along both the<br />

Carmel (Frechen et al. 2004) <strong>and</strong> Sharon (Engelmann<br />

et al. 2001; Frechen et al. 2001) coasts, with the <strong>for</strong>mer<br />

representing a much longer temporal sequence since


524 Zenobia Jacobs BOREAS<br />

MIS 6, including evidence of the MIS 5e high sea-level<br />

in the <strong>for</strong>m of beachrock. The deposits along the Sharon<br />

coast date to o65 kyr. These TL <strong>and</strong> IRSL ages are<br />

imprecise (10–20% relative errors), however, <strong>and</strong> are<br />

commonly not in agreement with TL showing significantly<br />

more scatter than IRSL. The IRSL ages are<br />

also systematically younger; this may be a result of<br />

anomalous fading <strong>for</strong> which no corrections were made,<br />

although preliminary tests suggested no fading (e.g.<br />

Frechen et al. 2004). The lack of precision prevents reliable<br />

calculation of rates of accumulation <strong>and</strong> separation<br />

of some units, making correlation with climate cycles<br />

difficult.<br />

Along the Sharon coast, Porat et al. (2004) used a<br />

SAAD IRSL procedure to obtain ages with improved<br />

precision (5–10% relative errors). They found that all<br />

units comprising the most western kurkar ridge were<br />

deposited during the last 65 kyr, which confirmed the<br />

earlier MAAD IRSL <strong>and</strong> TL age estimates (e.g. Engelmann<br />

et al. 2001; Frechen et al. 2002). They also<br />

found that the accumulation rates of the kurkar <strong>and</strong><br />

hamra were very different: 1–7 m/kyr <strong>and</strong> 0.1 m/kyr,<br />

respectively. Both types of deposit have relatively high<br />

carbonate contents, so the accumulation rate is the decisive<br />

factor in whether or not a deposit will result in the<br />

<strong>for</strong>mation of kurkar or hamra. All kurkar units were<br />

deposited during MIS 4, but slow deposition of hamra<br />

spanning MIS 3, 2 <strong>and</strong> 1 prevents any correlation to<br />

specific climatic events. Frechen et al. (2004) suggested<br />

that more climate variations are represented in the<br />

terrestrial deposits of the Carmel coast than in those<br />

on the Sharon coast. Although neither the kurkar<br />

nor hamra <strong>for</strong>mative episodes can be related to the<br />

glacial or interglacial periods of the Northern Hemisphere,<br />

they do seem to show some agreement with<br />

climate-related cycles identified in <strong>marine</strong> <strong>and</strong> terrestrial<br />

archives of the eastern Mediterranean (Frechen<br />

et al. 2004).<br />

Northwestern Europe<br />

<strong>Luminescence</strong> dating studies applied to late Holocene<br />

s<strong>and</strong> dunes in northwestern Europe have proliferated in<br />

recent times. Holocene dune fields are well developed<br />

along these coasts <strong>and</strong> constitute ‘soft’ coasts that are<br />

especially susceptible to the impact of predicted sealevel<br />

rise <strong>and</strong> increased storminess as a result of global<br />

warming. Underst<strong>and</strong>ing the Holocene evolution of<br />

dunes in relation to established patterns of environmental<br />

change can, there<strong>for</strong>e, provide valuable in<strong>for</strong>mation<br />

<strong>for</strong> effective management of fragile <strong>coastal</strong><br />

systems, including the protection of many low-lying<br />

areas in northwestern Europe.<br />

The time of <strong>for</strong>mation of these <strong>coastal</strong> dunefields has<br />

been documented in a number of articles based on<br />

SAAD IRSL dating of feldspar grains (e.g. Wilson et al.<br />

2001, 2004; Clarke et al. 2002; Clarke & Rendell 2006)<br />

or SAR OSL dating of quartz grains (e.g. Clemmensen<br />

et al. 2001a, b; Sommerville et al. 2003; Clemmensen &<br />

Murray 2006; Nielsen et al. 2006; Aagaard et al. 2007).<br />

It can be deduced from these luminescence <strong>chronologies</strong><br />

that dune <strong>for</strong>mation was episodic; periods of<br />

s<strong>and</strong> drift <strong>and</strong> dune <strong>for</strong>mation alternated with periods<br />

of stabilization <strong>and</strong> soil <strong>for</strong>mation. Many of these studies<br />

have suggested that dunefield development was initiated<br />

by cooler <strong>and</strong> stormy periods (i.e. increased<br />

storminess <strong>and</strong> storm surges), which may be a result of<br />

the southward displacement of polar water <strong>and</strong> sea ice<br />

in the North Atlantic Ocean, creating an increased<br />

thermal gradient that caused a number of deep cyclones<br />

to pass over northwestern Europe. This hypothesis is<br />

based, in part, on the recurrent identification of dune<br />

<strong>for</strong>mation during the Little Ice Age (AD 1350–1800) in<br />

different parts of northwestern Europe (e.g. Bailey et al.<br />

2001; Clemmensen et al. 2001b; Wilson et al. 2001,<br />

2004; Clarke et al. 2002; Clarke & Rendell 2006). The<br />

most recent phase of <strong>coastal</strong> dune-building in northwest<br />

Europe was also identified from luminescence<br />

dating, which suggests that it typically ended 100–200<br />

years ago (e.g. Clarke & Rendell 2006; Clemmensen &<br />

Murray 2006).<br />

The effects of global warming <strong>and</strong> ongoing relative<br />

sea-level change may also be assessed using luminescence<br />

dating techniques. Some studies have determined<br />

past changes in sea level from the construction <strong>and</strong><br />

timing of beach ridges (e.g. Or<strong>for</strong>d et al. 2003; Nielsen<br />

et al. 2006; Bjrnsen et al. 2007; Roberts & Plater 2007),<br />

whereas in other studies estuarine (e.g. Madsen et al.<br />

2005, 2007a), mud-flat (e.g. Madsen et al. 2007b; Mauz<br />

& Bungenstock 2007) <strong>and</strong> salt-marsh (e.g. Madsen et al.<br />

2007a) deposits have been investigated. The latter environments<br />

may be more sensitive indicators of ongoing<br />

sea-level change, as well as providing a means by<br />

which to evaluate sediment stability <strong>and</strong> budgeting in<br />

<strong>coastal</strong> estuarine <strong>and</strong> lagoonal settings. However, dating<br />

young <strong>sediments</strong> derived from such depositional<br />

environments necessitates careful investigation of the<br />

bleaching properties of the grains, the effect of changing<br />

burial depth on dose-rate estimates, <strong>and</strong> the difficulties<br />

inherent in determining water content. Madsen<br />

et al. (2005) presented a detailed luminescence study in<br />

which all these different aspects are described. Madsen<br />

et al. (2005, 2007a) also tested the accuracy of their assumptions<br />

using independent age estimates derived<br />

from 210 Pb <strong>and</strong> 137 Cs measurements. They demonstrated<br />

good agreement of OSL ages with 210 Pb <strong>and</strong><br />

137 Cs ages, from which they inferred that the <strong>sediments</strong><br />

have received sufficient sunlight exposure at deposition.<br />

But they also reported that the D e values obtained from<br />

the uppermost layers were very scattered, despite largesize<br />

aliquots being used to maximize the OSL output.<br />

Scatter in D e values <strong>for</strong> such large aliquots suggests either<br />

partial bleaching or significant mixing as a result of


BOREAS <strong>Luminescence</strong> <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> 525<br />

bioturbation. Neither scenario was explicitly examined<br />

<strong>and</strong> thus precludes a demonstration of accuracy <strong>and</strong><br />

precision <strong>for</strong> the ages obtained <strong>for</strong> the uppermost levels,<br />

as reflected in their 17% relative uncertainties. Ballarini<br />

et al. (2007) demonstrated that partial bleaching<br />

is prevalent in very young (o10 years) aeolian dune<br />

s<strong>and</strong>s on Texel Isl<strong>and</strong> in The Netherl<strong>and</strong>s, <strong>and</strong> that<br />

measurement of individual grains allowed detection of<br />

partial bleaching, based on analysis of D e distributions.<br />

There<strong>for</strong>e, combining dose-rate modelling with D e distribution<br />

analysis <strong>for</strong> small aliquots or single grains can<br />

further improve the prospects of dating these types of<br />

deposit. Importantly, these studies (e.g. Madsen et al.<br />

2005, 2007a, b; Mauz & Bungenstock 2007; Roberts &<br />

Plater 2007) demonstrate that OSL dating has great<br />

potential to be applied to ‘difficult’ deposits <strong>and</strong>, as<br />

such, make a significant contribution to addressing issues<br />

of real-time climate change <strong>and</strong> predictive modelling<br />

of future impacts.<br />

Oceans <strong>and</strong> seas<br />

About 70% of the Earth’s surface is covered by oceans<br />

<strong>and</strong> seas <strong>and</strong> much of our underst<strong>and</strong>ing of past environments<br />

<strong>and</strong> climate cycles hinges on knowledge<br />

gathered from analysis of continuous sediment records<br />

obtained from deep-sea cores. Application of luminescence<br />

dating to this environment has been limited,<br />

however. In the l<strong>and</strong>mark studies of Wintle & Huntley<br />

(1979a, b, 1980), TL dating was applied to two ocean<br />

cores collected from the North Pacific (TT28-14) <strong>and</strong><br />

North Atlantic (RC8-39) in water depths of 5079 m <strong>and</strong><br />

4330 m, respectively. They obtained TL ages ranging<br />

between 9.3 <strong>and</strong> 140 kyr using a MAAD, partial-bleach<br />

TL procedure applied to polymineral fine grains (4–11 m<br />

m). These TL ages were stratigraphically coherent, but<br />

the uncertainties were large, due mainly to inadequate<br />

exposure of <strong>sediments</strong> to sunlight prior to deposition,<br />

the moisture content, the presence of anomalous fading,<br />

<strong>and</strong> the need to make complicated time-dependent<br />

dose-rate corrections to allow <strong>for</strong> the presence of excess<br />

230 Th <strong>and</strong> 231 Pa in deep-ocean <strong>marine</strong> <strong>sediments</strong>.<br />

Following the ground-breaking work of Wintle <strong>and</strong><br />

Huntley (1979a), Berger et al. (1984) used known-age<br />

continental shelf material that did not suffer from<br />

excess 230 Th <strong>and</strong> 231 Pa, to develop a more accurate<br />

method of correcting <strong>for</strong> the light-insensitive residual<br />

dose. Problems of this type discouraged further application<br />

of luminescence dating to such <strong>sediments</strong>.<br />

It was more than two decades later, after the SAR<br />

procedure <strong>for</strong> measurement of single aliquots <strong>and</strong><br />

grains of quartz had been established, that dating of<br />

deep-sea <strong>sediments</strong> was re-visited. In part, this was because<br />

the most light-sensitive signals from quartz <strong>and</strong><br />

feldspar could now be exploited. Stokes et al. (2003)<br />

applied SAR OSL to single aliquots of silt-sized quartz<br />

to test the suitability of the OSL dating technique <strong>for</strong><br />

oceanic <strong>sediments</strong>, <strong>and</strong> they used thin source alpha<br />

spectrometry to examine the U <strong>and</strong> Th decay series <strong>for</strong><br />

time-dependent changes in the dose rate. They also<br />

provided updated dose-rate conversion factors <strong>for</strong> the<br />

major radionuclides in these decay chains. They dated<br />

nine samples from two independently dated sediment<br />

cores extracted from the Arabian Sea, <strong>and</strong> obtained<br />

OSL ages ranging between 7 kyr <strong>and</strong> 117 kyr with relative<br />

st<strong>and</strong>ard errors of 3–6%. With the exception of<br />

only one estimate, these ages agree well with the independent<br />

<strong>chronologies</strong>. Even though the silt-sized<br />

fraction used <strong>for</strong> dating in this study was dominated by<br />

aeolian input, <strong>and</strong> each aliquot was composed of several<br />

thous<strong>and</strong> grains, some aliquots exhibited a relatively<br />

large spread in D e values.<br />

Olley et al. (2004a) also dated quartz grains (but of<br />

60–70 mm diameter) from a core (Fr10/95-GC17) in the<br />

Indian Ocean, off the northwest Australian coast, in a<br />

water depth of 1093 m. They obtained ages ranging between<br />

1.8 kyr <strong>and</strong> 51 kyr, with typical undertainties of<br />

10%, that showed good agreement with 14 C ages. The<br />

OSL ages were obtained from individual grains, which<br />

allowed investigation of the D e distributions. They<br />

found that partial bleaching was a problem in two of<br />

their samples, but were able to calculate accurate ages<br />

using the minimum age model of Galbraith et al.<br />

(1999). These results serve as a warning that adequate<br />

pre-depositional bleaching of all traps giving rise to the<br />

OSL signal cannot be assumed in the <strong>marine</strong> environment,<br />

even in ideal situations such as this, where the<br />

core site lies beneath the pathway of aeolian dust<br />

transported from the Australian mainl<strong>and</strong>.<br />

Two luminescence dating studies have been conducted<br />

in the Arctic Ocean, where limited chronological,<br />

palaeoceanographic <strong>and</strong> palaeoclimatic<br />

in<strong>for</strong>mation exists due to the variable <strong>and</strong> erratic generation<br />

<strong>and</strong> preservation of <strong>for</strong>aminifera. Application<br />

of luminescence dating there<strong>for</strong>e has the potential to<br />

make a significant contribution to the construction of<br />

<strong>chronologies</strong> in this environment. In contrast to Olley<br />

et al. (2004a), Jakobsson et al. (2003) applied the multigrain<br />

SAR method to s<strong>and</strong>-sized (463 mm in diameter)<br />

quartz from the Lomonosov Ridge in the central Arctic<br />

Ocean. They obtained SAR ages of 110 kyr, consistent<br />

with MIS 5, <strong>for</strong> eight samples, with a ninth MIS<br />

5 sample yielding an age of 16511 kyr. Two samples<br />

collected at half the stratigraphic depth of the MIS 5<br />

samples gave ages of 115 kyr, which clearly overestimated<br />

their expected stratigraphic ages. They used<br />

high-resolution gamma spectrometry to check <strong>for</strong> unsupported<br />

230 Th, observing little excess in most samples;<br />

significant 226 Ra disequilibrium (compared to its<br />

parent 238 U) was reported <strong>for</strong> one sample. In a more<br />

comprehensive study, Berger (2006) determined ages<br />

<strong>for</strong> 9 core-top <strong>and</strong> 37 deeper sediment samples collected<br />

from 20 different cores, representing 19 different sites


526 Zenobia Jacobs BOREAS<br />

across the Arctic Ocean. Berger (2006) applied a<br />

MAAD IRSL <strong>and</strong> TL procedure to silt-sized (4–11 mm)<br />

grains extracted from these samples. The core-top<br />

samples were measured to test the TL <strong>and</strong> IRSL<br />

bleaching characteristics of the <strong>sediments</strong> <strong>and</strong> to assess<br />

the possible impact of partial bleaching on the ages obtained<br />

<strong>for</strong> older samples. Berger (2006) suggested that,<br />

based on the measurement of these core-top samples,<br />

bleaching of the traps responsible <strong>for</strong> both the TL <strong>and</strong><br />

IRSL signals is likely to be highly variable across the<br />

Arctic Ocean <strong>and</strong> will depend on whether the <strong>sediments</strong><br />

were deposited during a glacial or an interglacial period;<br />

he also noted the likelihood of some regional variation<br />

in the adequacy of sunlight exposure at the time<br />

of sediment deposition. Some of the core-top samples<br />

produced satisfactory results using a novel, shortbleach<br />

IRSL procedure, which may prove useful in future<br />

studies. Dating of the deeper <strong>sediments</strong> indicates<br />

the potential of TL <strong>and</strong> IRSL to extend beyond the effective<br />

age-range of 14 C dating, but currently their reliability<br />

is unknown. Berger (2006) also found that<br />

time-dependent dose rate corrections were less significant<br />

in the Arctic Ocean than further south.<br />

<strong>Luminescence</strong> dating of deep-sea <strong>sediments</strong> remains<br />

difficult <strong>and</strong> subject to assumptions about adequate<br />

bleaching of the luminescence signals, as well as possible<br />

time-dependent changes in the dose rate. As a consequence,<br />

such studies require significant extra input<br />

compared to luminescence dating of terrestrial deposits.<br />

But these studies have demonstrated the feasibility<br />

of OSL dating of sub<strong>marine</strong> <strong>sediments</strong>, at least in ‘ideal’<br />

environments such as those along the Australian <strong>and</strong><br />

Arabian coasts.<br />

Potential pitfalls<br />

One of the most significant aspects of luminescence dating<br />

is its intrinsically experimental character. Because of<br />

the variability in the luminescence properties of natural<br />

quartz <strong>and</strong> feldspar, <strong>and</strong> the variety of processes involved<br />

in sediment transport <strong>and</strong> deposition, the techniques<br />

applied to each set of samples require testing, <strong>and</strong><br />

suitable measurement conditions ought to be validated<br />

each time. In a large proportion of articles, this has been<br />

achieved to some extent, more so in those that appeared<br />

after 2000 than be<strong>for</strong>e, because of the relative ease of<br />

doing such tests with the SAR procedures.<br />

Signal resetting (bleaching) by sunlight<br />

In most articles, some assessment of the extent of sunlight<br />

bleaching has been made using either a modern<br />

analogue (e.g. Boyd et al. 2008) or controlled laboratory<br />

bleaching experiments. In many cases, however, it<br />

was uncertain whether the modern analogue was representative<br />

of all types of <strong>sediments</strong> <strong>and</strong> processes involved.<br />

An assessment of bleaching is especially<br />

important when the TL signal is utilized <strong>and</strong> the depositional<br />

process is something other than aeolian (e.g.<br />

beach, tsunami, storm <strong>and</strong> estuarine deposits). In a<br />

series of actualistic studies, Rink (1998), Rink & Pieper<br />

(2001) <strong>and</strong> Richardson (2001) explicitly measured the<br />

amount of natural ‘residual’ TL (NRTL) in a modern<br />

sample (i.e. the TL arising from unbleached traps in<br />

modern <strong>sediments</strong>). Rink (1998) showed that the<br />

NRTL in quartz grains collected from the surface of a<br />

beach exhibited a strong gradient as a function of its<br />

distance from the shoreline, while Rink & Pieper (2001)<br />

demonstrated that beach surfaces (i.e. the upper 1 cm of<br />

deposit) <strong>and</strong> dunes had significantly lower NRTL values<br />

than did <strong>sediments</strong> from the nearshore underwater<br />

zone (70 cm water depth). This outcome is not surprising,<br />

as the UV component of sunlight is strongly<br />

absorbed by sea water. Rink & Pieper (2001) also studied<br />

the NRTL of quartz grains in a modern storm<br />

berm, a tidal berm <strong>and</strong> the structures within them,<br />

which had been <strong>for</strong>med as a result of a hurricane along<br />

the Gulf of Mexico in Florida. High NRTL values were<br />

observed in all cases. The traps responsible <strong>for</strong> the lightsensitive<br />

signals in quartz <strong>and</strong> feldspar are significantly<br />

easier to empty than are the TL traps, requiring shorter<br />

sunlight exposure times. But adequate exposure is not<br />

necessarily true <strong>for</strong> all types of sediment, as was demonstrated<br />

by Richardson (2001) <strong>for</strong> a modern sample<br />

collected from an estuary. In this study, light-reading<br />

measurements indicated a three orders-of-magnitude<br />

reduction in light penetration in less than 1 m of water:<br />

this was a worse-case scenario, where the water was<br />

generally turbid <strong>and</strong> the <strong>sediments</strong> were muddy. Nonetheless,<br />

ages of 2500 years (IRSL <strong>and</strong> GLSL) <strong>and</strong><br />

11 000 years (TL) <strong>for</strong> modern <strong>sediments</strong> were obtained.<br />

Modern surface s<strong>and</strong> <strong>and</strong> intertidal beach s<strong>and</strong> yielded<br />

IRSL D e values consistent with a zero age. Although<br />

these bleaching studies broadly differentiate between<br />

depositional environments in which <strong>sediments</strong> are likely<br />

to be well bleached or only partially bleached, these<br />

generalizations may not apply to every site; it is prudent,<br />

there<strong>for</strong>e, to conduct tests at each study location<br />

to validate the efficacy of bleaching.<br />

D e distribution analysis<br />

A conspicuous scarcity in the literature is the application<br />

of small-aliquot <strong>and</strong> single-grain dating combined<br />

with dose-distribution analysis. Such studies are commonplace<br />

in applications to fluvial deposits (e.g. Olley<br />

et al. 2004b) <strong>and</strong> archaeological <strong>sediments</strong> (e.g. Jacobs<br />

et al. 2006, 2008a; David et al. 2007; Jacobs & Roberts<br />

2007) <strong>and</strong> ought to be applied routinely to certain types<br />

of <strong>coastal</strong> deposit. Only a limited number of studies report<br />

the use of small aliquots (Zazo et al. 2005) <strong>and</strong><br />

single grains (Jacobs et al. 2003a; Bateman et al. 2004;


BOREAS <strong>Luminescence</strong> <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> 527<br />

Olley et al. 2004a, b; Goodwin et al. 2006; Ballarini<br />

et al. 2007; Bostock et al. 2007; Carr et al. 2007; Brooke<br />

et al. 2008a, c) in aeolian, <strong>marine</strong> <strong>and</strong> estuarine <strong>sediments</strong>.<br />

No single-grain D e distribution data were presented<br />

or discussed <strong>for</strong> some of the studies (e.g.<br />

Goodwin et al. 2006; Bostock et al. 2007; Brooke et al.<br />

2008c). Bostock et al. (2007), however, mentioned that<br />

some samples required the minimum age model to estimate<br />

the burial dose; this suggests that these D e distributions<br />

probably showed some evidence of partial<br />

bleaching. Good agreement was obtained between<br />

the quartz OSL <strong>and</strong> 14 C results, suggesting that singlegrain<br />

analysis was advantageous in this estuarine<br />

environment.<br />

Zazo et al. (2005) reported results <strong>for</strong> one sample<br />

analysed using 4200 small aliquots (each composed of<br />

o10 grains) <strong>and</strong> 24 larger aliquots (50 grains). Even<br />

though the large aliquots are small compared to those<br />

measured at most laboratories, the difference in results<br />

between their small <strong>and</strong> large aliquots is striking. The<br />

D e values <strong>for</strong> the small aliquots ranged from 55 to<br />

240 Gy, compared to 130 to 400 Gy <strong>for</strong> the larger<br />

aliquots. Such a range cannot be explained by beta microdosimetry,<br />

but partial bleaching or post-depositional<br />

mixing cannot be excluded. In this situation,<br />

neither small nor large aliquots will necessarily yield the<br />

correct depositional age, <strong>and</strong> single grains ought to be<br />

measured. The calculated ages should, there<strong>for</strong>e, be<br />

viewed with caution.<br />

Olley et al. (2004b) measured single grains from a<br />

<strong>marine</strong> sediment sample collected from a 3500-year-old<br />

storm deposited beach ridge in Queensl<strong>and</strong>, Australia.<br />

The D e values, when plotted as a radial plot, indicated<br />

that the sample had been affected significantly by partial<br />

bleaching, but these authors were able to obtain<br />

excellent agreement with a 14 C age when the minimum<br />

age model was applied. They also constructed 96 synthetic<br />

aliquots, of which 48 contained 100 grains while<br />

the other 48 contained 10 grains. When they applied the<br />

minimum age model to these two data sets, they obtained<br />

ages that were too old by 32% <strong>for</strong> the 10-grain<br />

<strong>and</strong> by 57% <strong>for</strong> the 100-grain aliquots. This illustrates<br />

clearly that caution should be exercised when using<br />

multi-grain aliquots, even small aliquots, <strong>for</strong> <strong>sediments</strong><br />

that may have been poorly bleached; this may explain<br />

the anomalous results obtained by Zazo et al. (2005).<br />

Partial bleaching was also observed, using single grains,<br />

<strong>for</strong> two samples from a deep-sea ocean core off the<br />

coast of northwestern Australia (Olley et al. 2004a). In<br />

addition to analysing the D e distributions <strong>for</strong> these two<br />

samples, they also estimated the D e values associated<br />

with the fast-dominated signal from LM-OSL measurements<br />

(e.g. Yoshida et al. 2003); this resulted in a<br />

similar D e distribution. Applying the minimum age<br />

model to these two samples resulted in age estimates<br />

consistent with 14 C ages. Marine <strong>sediments</strong> have,<br />

there<strong>for</strong>e, been shown not to be immune to partial or<br />

heterogeneous bleaching. Accordingly, caution should<br />

be taken in general <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong>,<br />

<strong>and</strong> especially when dating Holocene samples, where<br />

the effects of incomplete bleaching at burial will be exacerbated<br />

by the comparatively small dose absorbed<br />

after deposition.<br />

Single-grain analysis was also applied to aeolian deposits<br />

by Ballarini et al. (2007), who measured single<br />

grains <strong>for</strong> two modern samples from Texel Isl<strong>and</strong> in<br />

The Netherl<strong>and</strong>s. One sample was known to be o10<br />

years old <strong>and</strong> the other had been deposited 1 year be<strong>for</strong>e<br />

collection. Single-aliquot data suggested that the<br />

<strong>for</strong>mer had been well bleached, whereas the D e <strong>for</strong> the<br />

latter sample was significantly overestimated <strong>and</strong> resulted<br />

in an age of 7424 years. Using a modified SAR<br />

approach specifically designed <strong>for</strong> young samples, Ballarini<br />

et al. (2007) were able to obtain ages similar to<br />

those obtained from the single aliquots when a mean D e<br />

value was used. The single-grain D e distributions of<br />

these samples, however, provided valuable in<strong>for</strong>mation<br />

that clearly indicated the presence of some grains <strong>for</strong><br />

which not even the fast component had been sufficiently<br />

bleached. The presence of such grains in a multigrain<br />

aliquot will result in age overestimation, just as it<br />

will when using the mean of the single-grain D e values.<br />

Partial bleaching, there<strong>for</strong>e, can also afflict to a measurable<br />

extent samples with an aeolian origin, <strong>and</strong> the<br />

extent of this bias will be exacerbated in age determination<br />

<strong>for</strong> young samples. At present, application of<br />

the minimum age model to such young samples is not<br />

straight<strong>for</strong>ward, but use of the mean <strong>for</strong> such distributions<br />

is not appropriate either; a revision of the minimum<br />

age model that does not log trans<strong>for</strong>m the D e<br />

values has, there<strong>for</strong>e, been developed (R. F. Galbraith,<br />

pers. comm. 2008).<br />

In South Africa, Jacobs et al. (2003a) calculated D e<br />

values <strong>for</strong> two aeolianite samples at sea level, <strong>and</strong> <strong>for</strong> an<br />

uncemented dune s<strong>and</strong> deposited inside a <strong>coastal</strong> cave,<br />

using large, single aliquots (1000 grains), single grains<br />

<strong>and</strong> synthetic aliquots (each of 100 grains). They obtained<br />

consistent results <strong>for</strong> all three aeolian units,<br />

when applying the central age model of Galbraith et al.<br />

(1999). The single-grain D e distribution, however, was<br />

wider than expected (‘overdispersed’), but similar in<br />

spread to laboratory-controlled (dose recovery) measurements<br />

of quartz grains given a known dose. Based<br />

on the appearance of the data on a radial plot, it can be<br />

suggested that such scatter in apparent D e results from<br />

natural variability in quartz OSL that is not fully accounted<br />

<strong>for</strong> in the SAR procedure. Microdosimetry<br />

variations in the beta dose to individual grains is a<br />

possibility, but is considered unlikely as similar results<br />

were obtained from both cemented <strong>and</strong> uncemented<br />

samples. Also along the southern Cape coast of South<br />

Africa, Bateman et al. (2004) <strong>and</strong> Carr et al. (2007)<br />

measured single grains <strong>for</strong> a subset of their samples.<br />

For one of the older samples, about 5 m below the


528 Zenobia Jacobs BOREAS<br />

surface, Bateman et al. (2004) observed much greater<br />

dispersion in the single-grain D e values <strong>and</strong> a significant<br />

shift of 35% in the mean D e <strong>for</strong> single grains compared<br />

to the multi-grain aliquots; unexpectedly, the<br />

latter were smaller, but no explanation was offered. For<br />

the sample near the top of the profile (o1 m below the<br />

surface) a similar D e value was obtained from single<br />

grains <strong>and</strong> multiple-grain aliquots, with some high <strong>and</strong><br />

low D e outliers. Being located just below the zone with<br />

root activity, bioturbation may explain some of these<br />

outliers. Bateman et al. (2004) did not use the singlegrain<br />

results to calculate their ages, so a re-assessment<br />

of the age of at least the lower sample (Shfd02007) is<br />

warranted.<br />

Carr et al. (2007) showed single-grain D e distributions<br />

<strong>for</strong> four samples from the Wilderness seaward<br />

dune cordon in South Africa. For two of the samples<br />

(KT2-6 <strong>and</strong> KT2-8), the D e distributions, when displayed<br />

in radial plots, appear to represent a single depositional<br />

event. But application of the central age<br />

model to the single-grain <strong>and</strong> single-aliquot D e values<br />

does not yield compatible results. The central D e value<br />

<strong>for</strong> the single grains from sample KT2-6 is 32% lower<br />

than that obtained from the single aliquots, whereas <strong>for</strong><br />

sample KT2-8 the central D e value <strong>for</strong> the single grains<br />

is 22% higher than the corresponding single-aliquot<br />

value (i.e. there is no systematic bias). This result is<br />

perplexing <strong>and</strong> requires further investigation, as the<br />

causes of this discrepancy may influence all the singlealiquot<br />

ages calculated <strong>for</strong> the dune cordon. Carr et al.<br />

(2007) also reported two samples (KT2-1 <strong>and</strong> KT2-10)<br />

<strong>for</strong> which the single-grain D e distributions are more<br />

overdispersed than expected <strong>for</strong> quartz grains that had<br />

been well bleached at burial <strong>and</strong> undisturbed thereafter<br />

(59% <strong>and</strong> 37%, respectively). The appearance of both<br />

radial plots suggests possible postdepositional mixing,<br />

in which case the application of the finite mixture model<br />

(Jacobs & Roberts 2007) may be able to resolve some of<br />

the discrete dose components <strong>and</strong>, thereby, result in<br />

more accurate results. Carr et al. (2007) suggested the<br />

possibility of beta microdosimetry variations as an explanation,<br />

but the effects would be expected to be similar<br />

throughout the seemingly homogeneous deposit.<br />

However, this may not be the case if the D e distributions<br />

are also affected by spatial variability in the degree<br />

of diagenetic alteration (e.g. Murray-Wallace et al.<br />

2001). The effects of beta-dose heterogeneity could be<br />

calculated explicitly (e.g. Jacobs et al. 2008b), which<br />

would allow the magnitude of any associated D e dispersion<br />

to be compared directly. Single-grain measurements<br />

can also be useful <strong>for</strong> older samples where significantly<br />

skewed results are obtained, because some proportion of<br />

the grains may be fully saturated. It has previously been<br />

shown that improved resolution can be achieved when<br />

saturated grains, as well as the ‘Class 3’ type grains<br />

of Yoshida et al. (2000), are identified <strong>and</strong> excluded prior<br />

to age determination (e.g. Jacobs et al. 2008b).<br />

These studies show that, even in aeolian environments,<br />

it cannot be safely assumed that all grains received<br />

the same burial dose. Although sufficient<br />

exposure to sunlight may not be a significant problem<br />

in many instances, beta microdosimetry variations <strong>and</strong><br />

postdepositional mixing cannot be excluded <strong>and</strong> these<br />

can lead to significantly different D e estimates from<br />

single grains <strong>and</strong> single multi-grain aliquots. A narrow<br />

D e distribution obtained from the measurement of<br />

small or large aliquots cannot, there<strong>for</strong>e, be used as a<br />

guarantee of accuracy. It is recommended that single<br />

grains should routinely be measured <strong>for</strong> at least some of<br />

the samples in each study. Only then can we hope to<br />

obtain greater accuracy <strong>and</strong> precision <strong>for</strong> luminescence<br />

age estimates to rival those of other methods, such as<br />

U-series. Such improvement would enable luminescence<br />

dating to contribute to breakthroughs in answering<br />

some of the more pressing global issues related to<br />

future <strong>coastal</strong> evolution <strong>and</strong> sea-level change.<br />

Environmental dose-rate determination<br />

The lack of accuracy <strong>and</strong> precision is, in some cases,<br />

also a result of inappropriate estimates of the environmental<br />

dose rate. In carbonate environments, disequilibria<br />

at or near the head of the U-series chain are<br />

not unusual. Use of neutron activation analysis or inductively<br />

coupled plasma mass spectrometry analysis,<br />

both of which measure the U <strong>and</strong> Th parent concentrations,<br />

can result in substantial under- or overestimates<br />

of the dose rate. Few studies have made<br />

detailed dose-rate investigations. But there were a few<br />

studies that were driven by a specific problem associated<br />

with a certain type of depositional environment.<br />

One particular example is deep-ocean <strong>sediments</strong> <strong>for</strong><br />

which time-dependent corrections need to be made <strong>for</strong><br />

excess 230 Th <strong>and</strong> 231 Pa (e.g. Wintle & Huntley 1980;<br />

Stokes et al. 2003). Other potentially problematic environments<br />

include estuaries, where sediment compaction<br />

<strong>and</strong> accompanying changes in the gamma radiation<br />

field can be problematic (e.g. Madsen et al. 2005), <strong>and</strong><br />

Holocene-age littoral <strong>sediments</strong> that show significant<br />

parent excess in the U-series (e.g. Z<strong>and</strong>er et al. 2007).<br />

Nathan & Mauz (2008) have also demonstrated the extent<br />

to which dose rates in carbonate-rich <strong>sediments</strong>,<br />

such as aeolianites <strong>and</strong> beachrock, may be influenced by<br />

the replacement of moisture in pore spaces by secondary<br />

carbonates. In some cases, the effects were negligible, but<br />

<strong>for</strong> others they were more substantial, indicating the<br />

need to assess the possible impact on a site-by-site basis.<br />

Many studies use integral counting methods, such as in<br />

situ gamma spectrometry, thick-source alpha counting<br />

<strong>and</strong> beta counting to estimate the dose rates; these<br />

methods are preferable to those that measure only the<br />

parent U <strong>and</strong> Th concentrations, although additional<br />

analyses are required to correct <strong>for</strong> any time-dependent<br />

changes in the dose rate.


BOREAS <strong>Luminescence</strong> <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> 529<br />

Future prospects<br />

So, how well are we doing with luminescence dating<br />

when applied to <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong>? The<br />

lack of precision compared to other dating methods<br />

continues to be a major drawback, but this can be improved<br />

substantially by making increased use of D e<br />

distribution analysis <strong>and</strong> more appropriate dose-rate<br />

estimations. At the younger end of the datable age<br />

range, luminescence dating has improved significantly,<br />

owing to the more widespread application of SAR OSL<br />

procedures <strong>for</strong> quartz. The improved accuracy <strong>and</strong><br />

precision of OSL ages <strong>for</strong> young <strong>coastal</strong> deposits now<br />

enable geomorphologists to quantify the tempo at<br />

which the most recently deposited barrier evolved (e.g.<br />

Goodwin et al. 2006). This was, until recently, restricted<br />

to deposits that contained sufficient organic materials<br />

<strong>for</strong> 14 C dating <strong>and</strong> to those <strong>sediments</strong> <strong>for</strong> which the<br />

<strong>marine</strong> reservoir effect was not detrimental. So, perhaps<br />

it can be said that the ability of OSL to date young<br />

<strong>coastal</strong> deposits has produced a revolution in <strong>coastal</strong><br />

geomorphology. Further back in time, the relatively<br />

early saturation of the fast component in quartz, <strong>and</strong><br />

the prevalence of anomalous fading in potassium-rich<br />

feldspar grains, remain the major challenges at the<br />

maximum limits of luminescence dating. Alternative<br />

dating procedures developed explicitly to extend the<br />

age range, such as exploitation of the 3101C TL isothermal<br />

signal (Choi et al. 2006), the componentspecific<br />

‘slow’ OSL signal (Singarayer & Bailey 2003;<br />

Rhodes et al. 2006) <strong>and</strong> the ‘recuperated’ OSL signal<br />

(Wang et al. 2006) are currently limited by a lack of<br />

precision or by the lack of testing on <strong>coastal</strong> <strong>sediments</strong>.<br />

Also, the latter pair of methods access OSL traps that<br />

are very hard to bleach; this effectively restricts these<br />

procedures to <strong>sediments</strong> that have been exposed to<br />

many hours or days of sunlight prior to deposition.<br />

There are, there<strong>for</strong>e, some further avenues to explore in<br />

luminescence dating of <strong>coastal</strong> <strong>sediments</strong>, with the<br />

main aim being to improve the accuracy <strong>and</strong> precision<br />

of age estimates.<br />

Conclusions<br />

A large body of readily accessible literature is available<br />

describing the application of luminescence dating<br />

methods to many types of <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> deposits<br />

<strong>and</strong> addressing a diverse range of local, regional <strong>and</strong><br />

global issues. Thus far, luminescence dating has played<br />

a valuable role in improving our underst<strong>and</strong>ing of local<br />

<strong>and</strong> regional issues of l<strong>and</strong><strong>for</strong>m evolution <strong>and</strong> sea-level<br />

change. Although large strides have been taken towards<br />

studying the recent past to predict future changes in<br />

climate (e.g. Goodwin et al. 2006; Brooke et al. 2008c),<br />

it has, to a large extent, not yet achieved the necessary<br />

levels of accuracy or precision to be generally incorporated<br />

into modelling studies of global climate<br />

change <strong>and</strong> predictions of future trends. Technological<br />

<strong>and</strong> methodological advances made over the past few<br />

years should enable luminescence dating to begin to<br />

rival more established methods, but this will require the<br />

execution of much larger-scale <strong>and</strong> more systematic<br />

studies, in which the distribution of D e values <strong>and</strong><br />

details of the dose rate are scrutinized on a sampleby-sample<br />

basis. Such refinements may facilitate luminescence<br />

ages of the required level of accuracy <strong>and</strong><br />

precision to be obtained in a variety of depositional<br />

contexts <strong>and</strong> geographic regions. Prospects also exist to<br />

extend the maximum limit of the dating technique, at<br />

least <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> that have been<br />

sufficiently bleached at the time of deposition.<br />

Acknowledgements. – I thank Ann Wintle, Brendan Brooke <strong>and</strong><br />

Jeong-Heon Choi <strong>for</strong> constructive reviews of the manuscript.<br />

References<br />

Aagaard, T., Or<strong>for</strong>d, J. & Murray, A. S. 2007: Environmental controls<br />

on <strong>coastal</strong> dune <strong>for</strong>mation; Skallingen Spit, Denmark. Geomorphology<br />

83, 29–47.<br />

Anderson, A., Roberts, R. G., Dickenson, W., Clark, G., Burley, D.,<br />

de Biran, A., Hope, G. & Nunn, P. 2006: Times of s<strong>and</strong>: Sedimentary<br />

history <strong>and</strong> archaeology at the Sigatoka Dunes, Fiji. Geoarcheology<br />

21, 131–154.<br />

Antonioli, F., Kershaw, S., Renda, P., Rust, D., Belluomini, G.,<br />

Cerasoli, M., Radtke, U. & Silenzi, S. 2006: Elevation of the last<br />

interglacial highst<strong>and</strong> in Sicily (Italy): A benchmark of <strong>coastal</strong> tectonics.<br />

Quaternary International 145, 3–18.<br />

Armitage, S. J., Botha, G. A., Duller, G. A. T., Wintle, A. G., Rebelo,<br />

L. P. & Momade, F. J. 2006: The <strong>for</strong>mation <strong>and</strong> evolution of the<br />

barrier isl<strong>and</strong>s of Inhaca <strong>and</strong> Bazaruto, Mozambique. Geomorphology<br />

82, 295–308.<br />

Bailey, S. D., Wintle, A. G., Duller, G. A. T. & Bristow, C. S. 2001:<br />

S<strong>and</strong> deposition during the last millennium at Aberffraw, Anglesey,<br />

North Wales as determined by OSL dating of quartz. Quaternary<br />

Science Reviews 20, 701–704.<br />

Balescu, S., Dumas, B., Gueremy, P., Lamothe, M., Lhenaff, R. &<br />

Raffy, J. 1997: Thermoluminescence dating tests of Pleistocene <strong>sediments</strong><br />

from uplifted shorelines along the southwest coastline of<br />

the Calabrian Peninsula (southern Italy). Palaeogeography, Palaeoclimatology,<br />

Palaeoecology 130, 25–41.<br />

Balescu, S., Packman, S. C. & Wintle, A. G. 1991: Chronological separation<br />

of interglacial raised beaches from northwestern Europe<br />

using thermoluminescence. Quaternary Research 35, 91–102.<br />

Balescu, S., Packman, S. C., Wintle, A. G. & Grun, ¨ R. 1992: Thermoluminescence<br />

dating of the Middle Pleistocene raised beach of<br />

Sangatte (Northern France). Quaternary Research 37, 390–396.<br />

Ballarini, M., Wallinga, J., Murray, A. S., Van Heteren, S., Oost, A.<br />

P., Bos, A. J. J. & Van Eijk, C. W. E. 2003: Optical dating of young<br />

<strong>coastal</strong> dunes on a decadal time scale. Quaternary Science Reviews<br />

22, 1011–1017.<br />

Ballarini, M., Wallinga, J., Wintle, A. G. & Bos, A. J. J. 2007: Analysis<br />

of equivalent-dose distributions <strong>for</strong> single grains of quartz<br />

from modern deposits. Quaternary Geochronology 2, 77–82.<br />

Banerjee, D., Hildebr<strong>and</strong>, A. N., Murray-Wallace, C. V., Bourman,<br />

R. P., Brooke, B. P. & Blair, M. 2003: SAR-OSL ages from the<br />

str<strong>and</strong>ed beach dune sequence in south-east South Australia. Quaternary<br />

Science Reviews 22, 1019–1026.<br />

Banerjee, D., Murray, A. S. & Foster, I. D. L. 2001: Scilly Isles, UK:<br />

Optical dating of a possible tsunami deposit from the 1755 Lisbon<br />

earthquake. Quaternary Science Reviews 20, 715–718.


530 Zenobia Jacobs BOREAS<br />

Baretto, A. M. F., Bezerra, F. H. R., Sugio, K., Tatumi, S. H., Yee,<br />

M., Paiva, R. P. & Munita, C. S. 2002: Late Pleistocene <strong>marine</strong><br />

terrace deposits in northeastern Brazil: Sea-level change <strong>and</strong> tectonic<br />

implications. Palaeogeography, Palaeoclimatology, Palaeoecology<br />

179, 57–69.<br />

Bateman, M. D. & Catt, J. A. 1996: An absolute chronology <strong>for</strong> the<br />

raised beach <strong>and</strong> associated deposits at Sewerby, East Yorkshire,<br />

Engl<strong>and</strong>. Journal of Quaternary Science 11, 389–395.<br />

Bateman, M. D. & Murton, J. B. 2006: The chronostratigraphy of<br />

Late Pleistocene glacial <strong>and</strong> periglacial aeolian activity in the Tuktoyaktuk<br />

Coastl<strong>and</strong>s, NWT, Canada. Quaternary Science Reviews<br />

25, 2552–2568.<br />

Bateman, M. D., Holmes, P. J., Carr, A. S., Horton, B. P. & Jaiswal,<br />

M. K. 2004: Aeolianite <strong>and</strong> barrier-dune construction during the<br />

last two glacial – interglacial cycles from the southern Cape coast,<br />

South Africa. Quaternary Science Reviews 23, 1681–1698.<br />

Belperio, A. P. 1995: The Quaternary. In Drexel, J. F. & Preiss, W. V.<br />

(eds.): The Geology of South Australia, 218–281. Geological Survey<br />

of South Australia, Bulletin.<br />

Berger, G. W. 2006: Trans-arctic-ocean tests of fine-silt luminescence<br />

sediment dating provide a basis <strong>for</strong> an additional geochronometer<br />

<strong>for</strong> this region. Quaternary Science Reviews 25, 2529–2551.<br />

Berger, G. W., Huntley, D. J. & Stipp, J. J. 1984: Thermoluminescence<br />

studies on a C-14-dated <strong>marine</strong> core. Canadian Journal<br />

of Earth Sciences 21, 1145–1150.<br />

Berger, G. W., Burke, R. M., Carver, G. A. & Easterbrook, D. J.<br />

1991: Test of thermoluminescence dating with <strong>coastal</strong> <strong>sediments</strong><br />

from northern Cali<strong>for</strong>nia. Chemical Geology 87, 21–37.<br />

Berger, G. W., Murray, A. S. & Havholm, K. G. 2003: Photonic dating<br />

of Holocene back-barrier <strong>coastal</strong> dunes, northern North Carolina,<br />

USA. Quaternary Science Reviews 22, 1043–1050.<br />

Bitinas, A., Damusyte, A., Hutt, ¨ G., Jaek, I. & Kabailiene, M. 2001:<br />

Application of the OSL dating <strong>for</strong> stratigraphic correlation of Late<br />

Weichselian <strong>and</strong> Holocene sediment in the Lithuanian maritime<br />

region. Quaternary Science Reviews 20, 767–772.<br />

Bjrnsen, M., Clemmensen, L. B., Murray, A. & Pedersen, K. 2007:<br />

New evidence of the Littorina transgressions in the Kattegat: Optically<br />

stimulated luminescence dating of a beach ridge system on<br />

Anholt, Denmark. Boreas 37, 157–168.<br />

Blum, M. D., Carter, A. E., Zayac, T. & Goble, R. J. 2002: Middle<br />

Holocene sea-level <strong>and</strong> evolution of the Gulf of Mexico Coast (US).<br />

Journal of Coastal Research 36, 65–80.<br />

Boomer, I. & Horton, B. P. 2006: Holocene relative sea-level movements<br />

along the North Norfolk coast, UK. Palaeogeography, Palaeoclimatology,<br />

Palaeoecology 230, 32–51.<br />

Bostock, H. C., Brooke, B. P., Ryan, D., Hancock, G., Pietsch, T.,<br />

Packett, B. & Harle, K. 2007: Holocene <strong>and</strong> modern sediment storage<br />

in the subtropical macrotidal Fitzroy River estuary, southeast<br />

Queensl<strong>and</strong>, Australia. Sedimentary Geology 201, 321–340.<br />

Bothner, M. H. & Johnson, N. M. 1969: Natural thermoluminescent<br />

dosimetry in Late Pleistocene pelagic <strong>sediments</strong>. Journal of Geophysical<br />

Research 74, 5331–5338.<br />

Bourman, R. P., Belperio, A. P., Murray-Wallace, C. V. & Cann, J.<br />

H. 1999: A last interglacial embayment fill at Normanville South<br />

Australia, <strong>and</strong> its neotectonic implications. Transactions of the<br />

Royal Society of South Australia 123, 1–15.<br />

Boyd, R., Ruming, K., Goodwin, I. D., S<strong>and</strong>strom, M. & Schroder-<br />

¨<br />

Adams, C. 2008: Highst<strong>and</strong> transport of <strong>coastal</strong> s<strong>and</strong> to the deep<br />

ocean: A case study from Fraser Isl<strong>and</strong>, southeast Australia. Geology<br />

36, 15–18.<br />

Bristow, C. S. & Pucillo, K. 2006: Quantifying rates of <strong>coastal</strong> progradation<br />

from sediment column using GPR <strong>and</strong> OSL: The Holocene<br />

fill of Guichen Bay, south-east South Australia. Sedimentology<br />

53, 769–788.<br />

Brooke, B. P., Lee, R., Cox, M., Olley, J. M. & Pietsch, T. 2008a:<br />

Rates of shoreline progradation during the last 1,770 years at<br />

Beachmere, southeastern Queensl<strong>and</strong>, Australia, based on OSL<br />

dating of beach ridges. Journal of Coastal Research 24, 640–648.<br />

Brooke, B. P., Preda, M., Lee, R., Cox, M., Olley, J. M., Pietsch, T. &<br />

Price, D. M. 2008b: Development, composition <strong>and</strong> age of indurated<br />

s<strong>and</strong> layers in the Late Quaternary <strong>coastal</strong> deposits of<br />

northern Moreton Bay, Queensl<strong>and</strong>. Australian Journal of Earth<br />

Sciences 55, 141–158.<br />

Brooke, B. P., Ryan, D., Pietsch, T., Olley, J. M., Douglas, G.,<br />

Packett, R., Radke, L. & Flood, P. 2008c: Influence of climate<br />

fluctuations <strong>and</strong> changes in catchment l<strong>and</strong> use on Late Holocene<br />

<strong>and</strong> modern beach-ridge sedimentation on a tropical macrotidal<br />

coast: Keppel Bay, Queensl<strong>and</strong>, Australia. Marine Geology 251,<br />

195–208.<br />

Brooke, B. P., Ryan, D., Radke, L., Pietsch, T., Olley, J. M., Douglas,<br />

G., Flood, P. & Packett, B. 2006: A 1500 Year Record of Coastal<br />

Sediment Accumulation Preserved in Beach Deposits at Keppel Bay,<br />

Queensl<strong>and</strong>, Australia, 48 pp. Cooperative Research Centre <strong>for</strong><br />

Coastal Zone, Estuary <strong>and</strong> Waterway Management.<br />

Brooke, B. P., Young, R. W., Bryant, E. A., Murray-Wallace, C. V. &<br />

Price, D. M. 1994: A Pleistocene origin <strong>for</strong> shore plat<strong>for</strong>ms along<br />

the northern Illawarra coast, New South Wales. Australian Geographer<br />

25, 178–185.<br />

Bruckner, ¨ H. 1986: Stratigraphy, evolution <strong>and</strong> age of Quaternary<br />

<strong>marine</strong> terraces in Morocco <strong>and</strong> Spain. Zeitschrift für Geomorphologie<br />

62, 83–101.<br />

Bryant, E. A. & Price, D. M. 1997: Late Pleistocene <strong>marine</strong> chronology<br />

of the Gippsl<strong>and</strong> Lakes region, Australia. Physical Geography<br />

18, 318–334.<br />

Bryant, E. A., Young, R. W. & Price, D. M. 1992: Evidence of tsunami<br />

sedimentation on the southeastern coast of Australia. Journal<br />

of Geology 100, 753–765.<br />

Bryant, E. A., Young, R. W. & Price, D. M. 1996: Tsunami as a major<br />

control on <strong>coastal</strong> evolution, southeastern Australia. Journal of<br />

Coastal Research 12, 831–840.<br />

Bryant, E. A., Young, R. W. & Price, D. M. 1997a: Late Pleistocene<br />

<strong>marine</strong> deposition <strong>and</strong> TL chronology of the New South<br />

Wales, Australian coastline. Zeitschrift für Geomorphologie 41,<br />

269–285.<br />

Bryant, E. A., Young, R. W., Price, D. M. & Short, S. A. 1990:<br />

Thermoluminescence <strong>and</strong> uranium – thorium <strong>chronologies</strong> of<br />

Pleistocene <strong>coastal</strong> l<strong>and</strong><strong>for</strong>ms of the Illawarra region, New South<br />

Wales. Australian Geographer 21, 101–112.<br />

Bryant, E. A., Young, R. W., Price, D. M. & Wheeler, D. J. 1997b:<br />

The impact of tsunami on the coastline of Jervis Bay, southeastern<br />

Australia. Physical Geography 18, 440–459.<br />

Buynevich, I. V., FitzGerald, D. M. & Goble, R. J. 2007: A 1500 yr<br />

record of North Atlantic storm activity based on optically dated<br />

relict beach scarps. Geology 35, 543–546.<br />

Carr, A. S., Bateman, M. D. & Holmes, P. J. 2007: Developing<br />

a 150 ka luminescence chronology <strong>for</strong> the barrier dunes of the<br />

southern Cape, South Africa. Quaternary Geochronology 2,<br />

110–116.<br />

Chen, Y. W., Chen, Y. G., Murray, A. S., Liu, T. K. & Lai, T. C.<br />

2003: <strong>Luminescence</strong> dating of neotectonic activity on the southwestern<br />

<strong>coastal</strong> plain, Taiwan. Quaternary Science Reviews 22,<br />

1223–1229.<br />

Chivas, A. R., Garcia, A., Van der Kaars, S., Couapel, M. J. J., Holt,<br />

S., Reeves, J. M., Wheeler, D. J., Switzer, A. D., Murray-Wallace,<br />

C. V., Banerjee, D., Price, D. M., Wang, S. X., Pearson, G., Edgar,<br />

N. T., Beau<strong>for</strong>t, L., De Deckker, P., Lawson, E. & Cecil, C. B.<br />

2001: Sea-level <strong>and</strong> environmental changes since the last interglacial<br />

in the Gulf of Carpentaria, Australia: An overview. Quaternary<br />

International 83–85, 19–46.<br />

Choi, J. H., Murray, A. S., Cheong, C. S., Hong, D. G. & Chang, H.<br />

W. 2003a: The resolution of stratigraphic inconsistency in the luminescence<br />

ages of <strong>marine</strong> terrace <strong>sediments</strong> from Korea. Quaternary<br />

Science Reviews 22, 1201–1206.<br />

Choi, J. H., Murray, A. S., Cheong, C. S., Hong, D. G. & Chang, H.<br />

W. 2006: Estimation of equivalent dose using quartz isothermal TL<br />

<strong>and</strong> the SAR procedure. Quaternary Geochronology 1, 101–108.<br />

Choi, J. H., Murray, A. S., Jain, M., Cheong, C. S. & Chang, H. W.<br />

2003b: <strong>Luminescence</strong> dating of well-sorted <strong>marine</strong> terrace <strong>sediments</strong><br />

on the southeastern coast of Korea. Quaternary Science Reviews<br />

22, 407–421.<br />

Clarke, M. L. & Rendell, H. M. 2006: Effects of storminess, s<strong>and</strong><br />

supply <strong>and</strong> the North Atlantic Oscillation on s<strong>and</strong> invasion <strong>and</strong><br />

<strong>coastal</strong> dune accretion in western Portugal. The Holocene 16,<br />

341–355.<br />

Clarke, M. L., Rendell, H. M., Pye, K., Tastet, J.-P., Pontee, N. I. &<br />

Masse, L. 1999: Evidence <strong>for</strong> the timing of dune development on


BOREAS <strong>Luminescence</strong> <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> 531<br />

the Aquitaine coast, southwest France. Zeitschrift für Geomorphologie<br />

116, 147–163.<br />

Clarke, M. L., Rendell, H. M., Tastet, J.-P., Clave, B. & Masse, L.<br />

2002: Late-Holocene s<strong>and</strong> invasion <strong>and</strong> North Atlantic storminess<br />

along the Aquitaine Coast, southwest France. The Holocene 12,<br />

231–238.<br />

Clemmensen, L. B. & Murray, A. S. 2006: The termination of the last<br />

major phase of aeolian s<strong>and</strong> movement, <strong>coastal</strong> dunefields, Denmark.<br />

Earth Surface Processes <strong>and</strong> L<strong>and</strong><strong>for</strong>ms 31, 203–216.<br />

Clemmensen, L. B., Murray, A. S., Beck, J. H. & Clausen, A. 2001a:<br />

Large-scale aeolian s<strong>and</strong> movement on the west coast of Jutl<strong>and</strong>,<br />

Denmark in late Subboreal to early Subatlantic time – a record of<br />

climate change or cultural impact? GFF 123, 193–203.<br />

Clemmensen, L. B., Pedersen, K., Murray, A. S. & Heinemeier, J.<br />

2006: A 7000-year record of <strong>coastal</strong> evolution, Vejers, SW Jutl<strong>and</strong>,<br />

Denmark. Bulletin of the Geological Society of Denmark 53, 1–22.<br />

Clemmensen, L. B., Pye, K., Murray, A. S. & Heinemeier, J. 2001b:<br />

Sedimentology, stratigraphy <strong>and</strong> l<strong>and</strong>scape evolution of a Holocene<br />

<strong>coastal</strong> dune system, Lodbjerg, NW Jutl<strong>and</strong>, Denmark. Sedimentology<br />

48, 3–27.<br />

Cooper, A. F. & Kostro, F. 2006: A tectonically uplifted <strong>marine</strong><br />

shoreline deposit, Knights Point, Westl<strong>and</strong>, New Zeal<strong>and</strong>. New<br />

Zeal<strong>and</strong> Journal of Geology <strong>and</strong> Geophysics 49, 203–216.<br />

Coutard, S., Lautridou, J. P., Rhodes, E. J. & Clet, M. 2006: Tectonic,<br />

eustatic <strong>and</strong> climatic significance of raised beaches of Val de Saire,<br />

Cotentin, Norm<strong>and</strong>y, France. Quaternary Science Reviews 25,<br />

595–611.<br />

David, B., Roberts, R. G., Magee, J., Mialanes, J., Turney, C., Bird,<br />

M., White, C., Fifield, L. K. & Tibby, J. 2007: Sediment mixing at<br />

Nonda Rock: Investigations of stratigraphic integrity at an early<br />

archaeological site in northern Australia <strong>and</strong> implications <strong>for</strong> the<br />

human colonisation of the continent. Journal of Quaternary Science<br />

22, 449–479.<br />

Dawson, A. G., Hindson, R., Andrade, C., Freitas, C., Parish, R. &<br />

Bateman, M. 1995: Tsunami sedimentation association with the<br />

Lisbon earthquake of 1 November AD1755: Boca do Rio, Algarve,<br />

Portugal. The Holocene 5, 209–215.<br />

De Carvalho, G. S., Granja, H. M., Loureiro, E. & Henriques, R.<br />

2006: Late Pleistocene <strong>and</strong> Holocene environmental changes in the<br />

<strong>coastal</strong> zone of northwestern Portugal. Journal of Quaternary Science<br />

21, 859–877.<br />

Duller, G. A. T. 1991: Equivalent dose determination using single<br />

aliquots. Nuclear Tracks <strong>and</strong> Radiation Measurements 18, 371–378.<br />

Duller, G. A. T. 1994: <strong>Luminescence</strong> dating of <strong>sediments</strong> using<br />

single aliquots: New procedures. Quaternary Science Reviews 13,<br />

149–156.<br />

Duller, G. A. T. 1995: <strong>Luminescence</strong> dating using single aliquots:<br />

Methods <strong>and</strong> applications. Radiation Measurements 24, 217–226.<br />

Duller, G. A. T. 1996: The age of the Koputaroa dunes, southwest<br />

North Isl<strong>and</strong>, New Zeal<strong>and</strong>. Palaeogeography, Palaeoclimatology,<br />

Palaeoecology 121, 105–114.<br />

El-Asmar, H. M. & Wood, P. 2000: Quaternary shoreline development:<br />

The northwestern coast of Egypt. Quaternary Science Reviews<br />

19, 1137–1149.<br />

Engelmann, A., Neber, A., Frechen, M., Boenigk, W. & Ronen, A.<br />

2001: <strong>Luminescence</strong> chronology of Upper Pleistocene <strong>and</strong> Holocene<br />

aeolianites from Netanya South – Sharon Coastal Plain, Israel.<br />

Quaternary Science Reviews 20, 799–804.<br />

Feathers, J. K. 2002: <strong>Luminescence</strong> dating in less than ideal conditions:<br />

Case studies from Klasies River main site <strong>and</strong> Duinefontein,<br />

South Africa. Journal of Archaeological Science 29, 177–194.<br />

Feathers, J. K. & Bush, D. A. 2000: <strong>Luminescence</strong> dating of Middle<br />

Stone Age deposits at Die Kelders. Journal of Human Evolution 38,<br />

91–119.<br />

Federici, P. R. & Pappalardo, M. 2006: Evidence of Marine Isotope<br />

Stage 5.5 highst<strong>and</strong> in Liguria (Italy) <strong>and</strong> its tectonic significance.<br />

Quaternary International 145, 68–77.<br />

Forman, S. L., Wintle, A. G., Thorleifson, L. H. & Wyatt, P. H. 1987:<br />

Thermoluminescence properties <strong>and</strong> age estimates <strong>for</strong> Quaternary<br />

raised <strong>marine</strong> <strong>sediments</strong>, Hudson Bay Lowl<strong>and</strong>, Canada. Canadian<br />

Journal of Earth Sciences 24, 2405–2411.<br />

Frechen, M., Dermann, B., Boenigk, W. & Ronen, A. 2001:<br />

<strong>Luminescence</strong> chronology of aeolianites from the section at Givat<br />

Olga – Coastal Plain of Israel. Quaternary Science Reviews 20,<br />

805–809.<br />

Frechen, M., Neber, A., Dermann, B., Tsatskin, A., Boenigk, W. &<br />

Ronen, A. 2002: Chronostratigraphy of aeolianites from the Sharon<br />

Coastal Plain of Israel. Quaternary International 89, 31–44.<br />

Frechen, M., Neber, A., Tsatskin, A., Boenigk, W. & Ronen, A. 2004:<br />

Chronology of Pleistocene sedimentary cycles in the Carmel<br />

Coastal Plain of Israel. Quaternary International 121, 41–52.<br />

Fumanal, M. P. 1995: Pleistocene dune systems in the Valencian Betic<br />

cliffs (Spain). INQUA Subcommission on Mediterranean <strong>and</strong> Black<br />

Sea Shorelines Newsletter 17, 32–38.<br />

Galbraith, R. F., Roberts, R. G., Laslett, G. M., Yoshida, H. & Olley,<br />

J. M. 1999: Optical dating of single <strong>and</strong> multiple grains of<br />

quartz from Jinmium rock shelter, northern Australia: Part I.<br />

Experimental design <strong>and</strong> statistical models. Archaeometry 41,<br />

339–364.<br />

Giannini, P. C. F., Sawakulchi, A. O., Martinho, C. T. & Tatumi, S.<br />

H. 2007: Eolian depositional episodes controlled by Late Quaternary<br />

relative sea level changes in the Imbituba – Laguna coast<br />

(southern Brazil). Marine Geology 237, 143–168.<br />

Goodwin, I. D., Stables, M. A. & Olley, J. M. 2006: Wave climate,<br />

s<strong>and</strong> budget <strong>and</strong> shoreline alignment evolution of the Iluka –<br />

Woody Bay s<strong>and</strong> barrier, northern New South Wales, Australia,<br />

since 3000 yr BP. Marine Geology 226, 127–144.<br />

Goudie, A. S., Colls, A., Stokes, S., Parker, A., White, K. & Al-Farraj,<br />

A. 2000: Latest Pleistocene <strong>and</strong> Holocene dune construction at<br />

the north-eastern edge of the Rub Al Khali, United Arab Emirates.<br />

Sedimentology 47, 1011–1021.<br />

Hall, A. M., Gordon, J. E., Whittington, G., Duller, G. A. T. &<br />

Heijnis, H. 2002: Sedimentology, palaeoecology <strong>and</strong> geochronology<br />

of Marine Isotope Stage 5 deposits on the Shetl<strong>and</strong> Isl<strong>and</strong>s,<br />

Scotl<strong>and</strong>. Journal of Quaternary Science 17, 51–67.<br />

Hall, A. M., Hansom, J. D., Williams, D. M. & Jarvis, J. 2006: Distribution,<br />

geomorphology <strong>and</strong> lithofacies of cliff-top storm deposits:<br />

Examples from the high-energy coasts of Scotl<strong>and</strong> <strong>and</strong> Irel<strong>and</strong>.<br />

Marine Geology 232, 133–155.<br />

Hansom, J. D. & Hall, A. M. 2008: Magnitude <strong>and</strong> frequency of extra-tropical<br />

North Atlantic cyclones: A chronology from cliff-top<br />

storm deposits. Quaternary International, doi: 10.1016/<br />

j.quaint.2007.11.010.<br />

Havholm, K. G., Ames, D. V., Whittecar, G. R., Wenell, B. A., Riggs,<br />

S. R., Jol, H. M., Berger, G. W. & Holmes, M. A. 2004: Stratigraphy<br />

of back-barrier <strong>coastal</strong> dunes, northern North Carolina <strong>and</strong><br />

Southern Virginia. Journal of Coastal Research 20, 980–999.<br />

Henshilwood, C. S., d’Errico, F., Vanhaeren, M., Van Niekerk, K. &<br />

Jacobs, Z. 2004: Middle Stone Age shell beads from South Africa.<br />

Science 304, 404.<br />

Henshilwood, C. S., d’Errico, F., Yates, R., Jacobs, Z., Tribolo, C.,<br />

Duller, G. A. T., Mercier, N., Sealy, J. C., Valladas, H., Watts, I. &<br />

Wintle, A. G. 2002: Emergence of modern human behaviour: Middle<br />

Stone Age engravings from South Africa. Science 295,<br />

1278–1280.<br />

Hong, D. G., Choi, M. S., Han, J.-H. & Cheong, C. S. 2003: Determination<br />

of sedimentation rate of a recently deposited tidal flat,<br />

western coast of Korea, using IRSL dating. Quaternary Science<br />

Reviews 22, 1185–1190.<br />

Huntley, D. J. & Clague, J. J. 1996: Optical dating of tsunami-laid<br />

s<strong>and</strong>s. Quaternary Research 46, 127–140.<br />

Huntley, D. J. & Johnson, N. M. 1976: Thermoluminescence as a<br />

potential means of dating siliceous ocean <strong>sediments</strong>. Canadian<br />

Journal of Earth Sciences 13, 593–596.<br />

Huntley, D. J. & Lamothe, M. 2001: Ubiquity of anomalous<br />

fading in K-feldspars <strong>and</strong> the measurement <strong>and</strong> correction <strong>for</strong><br />

it in optical dating. Canadian Journal of Earth Sciences 38,<br />

1093–1106.<br />

Huntley, D. J. & Prescott, J. R. 2001: Improved methodology <strong>and</strong><br />

new thermoluminescence ages <strong>for</strong> the dune sequence in south-east<br />

South Australia. Quaternary Science Reviews 20, 687–699.<br />

Huntley, D. J., Berger, G. W., Divigalpitiya, W. M. R. & Brown, T.<br />

A. 1983: Thermoluminescence dating of <strong>sediments</strong>. PACT 9,<br />

607–618.<br />

Huntley, D. J., Godfrey-Smith, D. I. & Thewalt, M. L. W. 1985a:<br />

Optical dating of <strong>sediments</strong>. Nature 313, 105–107.


532 Zenobia Jacobs BOREAS<br />

Huntley, D. J., Hutton, J. T. & Prescott, J. R. 1985b: South Australian<br />

s<strong>and</strong> dunes – a TL sediment test sequence. Preliminary results.<br />

Nuclear Tracks <strong>and</strong> Radiation Measurements 10, 757–758.<br />

Huntley, D. J., Hutton, J. T. & Prescott, J. R. 1993a: The str<strong>and</strong>ed<br />

beach dune sequence of south-east South Australia: A test of thermoluminescence<br />

dating, 0–800 ka. Quaternary Science Reviews 12,<br />

1–20.<br />

Huntley, D. J., Hutton, J. T. & Prescott, J. R. 1993b: Optical dating<br />

using inclusions within quartz grains. Geology 21, 1087–1090.<br />

Huntley, D. J., Hutton, J. T. & Prescott, J. R. 1994a: Further thermoluminescence<br />

dates from the dune sequence in the southeast of<br />

South Australia. Quaternary Science Reviews 13, 201–207.<br />

Huntley, D. J., Lian, O. B., Hu, J. & Prescott, J. R. 1994b: Test of<br />

luminescence dating making use of palaeomagnetic reversals. Ancient<br />

TL 12, 28–30.<br />

Huntley, D. J., Short, M. A. & Dunphy, K. 1996: Deep traps in quartz<br />

<strong>and</strong> their use <strong>for</strong> optical dating. Canadian Journal of Physics 74,<br />

81–91.<br />

Idnurm, M. & Cook, P. J. 1980: Palaeomagnetism of beach ridges in<br />

South Australia <strong>and</strong> the Milankovitch theory of ice ages. Nature<br />

286, 699–702.<br />

Illenberger, W. K. 1996: The geomorphic evolution of the Wilderness<br />

dune cordons, South Africa. Quaternary International 33, 11–20.<br />

Jacobs, Z. & Roberts, D. L. In press: Last interglacial age <strong>for</strong> the<br />

Nahoon fossil human footprints, southeast coast of South Africa.<br />

Quaternary Geochronology.<br />

Jacobs, Z. & Roberts, R. G. 2007: Advances in optically stimulated<br />

luminescence dating of individual grains of quartz from archaeological<br />

deposits. Evolutionary Anthropology 16, 210–223.<br />

Jacobs, Z., Duller, G. A. T. & Wintle, A. G. 2003a: Optical dating of<br />

dune s<strong>and</strong> from Blombos Cave, South Africa: II-single grain data.<br />

Journal of Human Evolution 44, 613–625.<br />

Jacobs, Z., Duller, G. A. T., Wintle, A. G. & Henshilwood, C. S.<br />

2006: Extending the chronology of deposits at Blombos Cave,<br />

South Africa, back to 140 ka using optical dating of single <strong>and</strong><br />

multiple grains of quartz. Journal of Human Evolution 51, 255–273.<br />

Jacobs, Z., Wintle, A. G. & Duller, G. A. T. 2003b: Optical dating of<br />

dune s<strong>and</strong> from Blombos Cave, South Africa: I – multiple grain<br />

data. Journal of Human Evolution 44, 597–610.<br />

Jacobs, Z., Wintle, A. G., Duller, G. A. T., Roberts, R. G. & Wadley,<br />

L. 2008a: New ages <strong>for</strong> the post-Howiesons Poort, late <strong>and</strong> final<br />

Middle Stone Age at Sibudu, South Africa. Journal of Archaeological<br />

Science 35, 1790–1807.<br />

Jacobs, Z., Wintle, A. G., Roberts, R. G. & Duller, G. A. T. 2008b:<br />

Equivalent dose distributions from single grains of quartz at Sibudu,<br />

South Africa: Context, causes <strong>and</strong> consequences <strong>for</strong> optical<br />

dating of archaeological deposits. Journal of Archaeological Science<br />

35, 1808–1820.<br />

Jain, M., Choi, J. H. & Thomas, P. J. 2008: The ultrafast OSL component<br />

in quartz: Origins <strong>and</strong> implications. Radiation Measurements<br />

43, 709–714.<br />

Jain, M., Murray, A. S. & Btter-Jensen, L. 2003: Characterisation of<br />

blue-light stimulated luminescence components in different quartz<br />

samples: Implications <strong>for</strong> dose measurements. Radiation Measurements<br />

37, 441–449.<br />

Jakobsson, M., Backman, J., Murray, A. S. & Lvlie, R. 2003: Optically<br />

stimulated luminescence dating support central Arctic Ocean<br />

cm-scale sedimentation rates. Geochemistry, Geophysics, Geosystems<br />

4, article 1016.<br />

Jungner, H., Korjonen, K., Heikkinen, O. & Wiedermann, A. M. 2001:<br />

<strong>Luminescence</strong> <strong>and</strong> radiocarbon dating of a dune series at Cape Kiw<strong>and</strong>a,<br />

Oregon, USA. Quaternary Science Reviews 20, 811–814.<br />

Kamaludin, B. H., Nakamura, T., Price, D. M., Woodroffe, C. D. &<br />

Fuji, S. 1993: Radiocarbon <strong>and</strong> thermoluminescence dating of the<br />

Old Alluvium from a <strong>coastal</strong> site in Perak, Malaysia. Sedimentary<br />

Geology 83, 199–210.<br />

Kaufman, D. S., Manley, W. F., Wolfe, A. P., Hu, F. S., Preece, S. J.,<br />

Westgate, J. A. & Forman, S. L. 2001: The last interglacial to glacial<br />

transition, Togiak Bay, southwestern Alaska. Quaternary Research<br />

55, 190–202.<br />

Kendrick, G. W., Wyrwoll, K. H. & Szabo, B. J. 1991: Pliocene –<br />

Pleistocene <strong>coastal</strong> events <strong>and</strong> history along the western margin of<br />

Australia. Quaternary Science Reviews 10, 419–439.<br />

Kennedy, D. M., Tannock, K. L., Crozier, M. J. & Rieser, U. 2007:<br />

Boulders of MIS 5 age deposited by a tsunami on the coast of Otago,<br />

New Zeal<strong>and</strong>. Sedimentary Geology 200, 222–231.<br />

Knight, J., Or<strong>for</strong>d, J., Wilson, P., Wintle, A. G. & Braley, S. M. 1998:<br />

Facies, age <strong>and</strong> controls on recent <strong>coastal</strong> s<strong>and</strong> dune evolution in<br />

North Norfolk, Eastern Engl<strong>and</strong>. Journal of Coastal Research 26,<br />

154–161.<br />

Kortekaas, M., Murray, A. S., S<strong>and</strong>gren, P. & Bjorck, ¨ S. 2007: OSL<br />

chronology <strong>for</strong> a sediment core from the southern Baltic Sea: A<br />

continuous sedimentation record since deglaciation. Quaternary<br />

Geochronology 2, 95–101.<br />

Lamothe, M. & Auclair, M. 1999: A solution to anomalous fading<br />

<strong>and</strong> age shortfalls in optical dating of feldspar minerals. Earth <strong>and</strong><br />

Planetary Science Letters 171, 319–323.<br />

Lees, B. G., Hayne, M. & Price, D. M. 1993: Marine transgression<br />

<strong>and</strong> dune initiation on western Cape York, northern Australia.<br />

Marine Geology 114, 81–89.<br />

Lees, B. G., Lu, Y. C. & Head, J. 1990: Reconnaissance thermoluminescence<br />

dating of northern Australian <strong>coastal</strong> dune systems.<br />

Quaternary Research 34, 169–185.<br />

Lees, B. G., Lu, Y. C. & Price, D. M. 1992: Thermoluminescence<br />

dating of dunes at Cape St. Lambert, East Kimberleys, northwestern<br />

Australia. Marine Geology 106, 131–139.<br />

Lees, B. G., Stanner, J., Price, D. M. & Lu, Y. 1995: Thermoluminescence<br />

dating of dune podzols at Cape Arnhem, northern<br />

Australia. Marine Geology 129, 63–75.<br />

Litchfield, N. J. & Lian, O. B. 2004: <strong>Luminescence</strong> age estimates of<br />

Pleistocene <strong>marine</strong> terrace <strong>and</strong> alluvial fan <strong>sediments</strong> associated<br />

with tectonic activity along <strong>coastal</strong> Otago, New Zeal<strong>and</strong>. New<br />

Zeal<strong>and</strong> Journal of Geology <strong>and</strong> Geophysics 47, 29–37.<br />

Lopez, G. I. & Rink, W. J. 2007: Characteristics of the burial environment<br />

related to quartz SAR-OSL dating at St. Vincent Isl<strong>and</strong>,<br />

NW Florida, USA. Quaternary Geochronology 2, 65–70.<br />

Lopez, G. I. & Rink, W. J. 2008: New quartz optically stimulated luminescence<br />

ages <strong>for</strong> beach ridges on the St. Vincent Isl<strong>and</strong> Holocene<br />

str<strong>and</strong> plain, Florida, United States. Journal of Coastal<br />

Research 24, 49–62.<br />

Madsen, A. T., Murray, A. S., Andersen, T. J. & Pejrup, M. 2007a:<br />

Temporal changes of accretion rates on an estuarine salt marsh<br />

during the late Holocene – reflection of local sea level changes? The<br />

Wadden Sea, Denmark. Marine Geology 242, 221–233.<br />

Madsen, A. T., Murray, A. S., Andersen, T. J. & Pejrup, M. 2007b:<br />

Optical dating of young tidal <strong>sediments</strong> in the Danish Wadden Sea.<br />

Quaternary Geochronology 2, 89–94.<br />

Madsen, A. T., Murray, A. S., Andersen, T. J., Pejrup, M. & Breuning-Madsen,<br />

H. 2005: Optically stimulated luminescence dating of<br />

young estuarine <strong>sediments</strong>: A comparison with Pb-210 <strong>and</strong> Cs-137<br />

dating. Marine Geology 214, 251–268.<br />

Mallinson, D., Burdette, K., Mahan, S. & Brook, G. 2008: Optically<br />

stimulated luminescence age controls on late Pleistocene <strong>and</strong> Holocene<br />

<strong>coastal</strong> lithosomes, North Carolina, USA. Quaternary Research<br />

69, 97–109.<br />

Marean, C. W., Bar-Matthews, M., Bernatchez, J., Fisher, E., Goldberg,<br />

P., Herries, A. I. R., Jacobs, Z., Jerardino, A., Karkanas, P.,<br />

Minichillo, T., Nilssen, P. J., Thompson, E., Watts, I. & Williams,<br />

H. M. 2007: Early human use of <strong>marine</strong> resources <strong>and</strong> pigment<br />

in South Africa during the Middle Pleistocene. Nature 449,<br />

905–908.<br />

Mauz, B. 1999: Late Pleistocene records of littoral processes at the<br />

Tyrrhenian Coast (Central Italy); depositional environments <strong>and</strong><br />

luminescence chronology. Quaternary Science Reviews 18,<br />

1173–1184.<br />

Mauz, B. & Bungenstock, F. 2007: How to reconstruct trends of late<br />

Holocene relative sea level: A new approach using tidal flat clastic<br />

<strong>sediments</strong> <strong>and</strong> optical dating. Marine Geology 237, 225–237.<br />

Mauz, B. & Hassler, U. 2000: <strong>Luminescence</strong> chronology of Late<br />

Pleistocene raised beaches in southern Italy: New data of relative<br />

sea-level changes. Marine Geology 170, 187–203.<br />

Mauz, B., Buccheri, G., Zoller, ¨ L. & Greco, A. 1997: Middle to Upper<br />

Pleistocene morphostructural evolution of the NW-coast of Sicily:<br />

Thermoluminescence dating <strong>and</strong> palaeontological – stratigraphical<br />

evaluations of littoral deposits. Palaeogeography, Palaeoclimatology,<br />

Palaeoecology 128, 269–285.


BOREAS <strong>Luminescence</strong> <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> 533<br />

Murari, M. K., Achyuthan, H. & Singhvi, A. K. 2007: <strong>Luminescence</strong><br />

studies on the <strong>sediments</strong> laid down by the December 2004 tsunami<br />

event: Prospects <strong>for</strong> the dating of palaeo tsunamis <strong>and</strong> <strong>for</strong> the estimation<br />

of sediment fluxes. Current Science 92, 367–371.<br />

Murray, A. S. & Clemmensen, L. B. 2001: <strong>Luminescence</strong> dating of<br />

Holocene aeolian s<strong>and</strong> movement, Thy, Denmark. Quaternary<br />

Science Reviews 20, 751–754.<br />

Murray, A. S. & Funder, S. 2003: Optically stimulated luminescence<br />

dating of a Danish Eemian <strong>coastal</strong> <strong>marine</strong> deposit: A test of accuracy.<br />

Quaternary Science Reviews 22, 1177–1183.<br />

Murray, A. S. & Wintle, A. G. 2000: <strong>Luminescence</strong> dating of quartz<br />

using an improved single-aliquot regenerative-dose protocol. Radiation<br />

Measurements 32, 57–73.<br />

Murray-Wallace, C. V., Banerjee, D., Bourman, R. P., Olley, J. M. &<br />

Brooke, B. P. 2002a: Optically stimulated luminescence dating of<br />

Holocene relict <strong>for</strong>edunes, Guichen Bay, South Australia. Quaternary<br />

Science Reviews 21, 1077–1086.<br />

Murray-Wallace, C. V., Belperio, A. P., Bourman, R. P., Cann, J. H.<br />

& Price, D. M. 1999: Facies architecture of a last interglacial<br />

barrier: A model <strong>for</strong> Quaternary barrier development from the<br />

Coorong to Mount Gambier Coastal Plain, southeastern Australia.<br />

Marine Geology 158, 177–195.<br />

Murray-Wallace, C. V., Belperio, A. P., Cann, J. H., Huntley, D. J. &<br />

Prescott, J. R. 1996: Late Quaternary uplift history, Mount<br />

Gambier region, South Australia. Zeitschrift für Geomorphologie<br />

106, 41–56.<br />

Murray-Wallace, C. V., Brooke, B. P., Cann, J. H., Belperio, A. P. &<br />

Bourman, R. P. 2001: Whole-rock aminostratigraphy of the Coorong<br />

<strong>coastal</strong> plain, South Australia: Towards a 1 million year record<br />

of sea-level highst<strong>and</strong>s. Journal of the Geological Society,<br />

London 158, 111–124.<br />

Murray-Wallace, C. V., Jones, B. G., Nghi, T., Price, D. M., Van<br />

Vinh, V., Tinh, T. N. & Nanson, G. C. 2002b: Thermoluminescence<br />

ages <strong>for</strong> a reworked <strong>coastal</strong> barrier, southeastern Vietnam: A<br />

preliminary report. Journal of Asian Earth Sciences 20, 535–548.<br />

Nathan, R. P. & Mauz, B. 2008: On the dose-rate estimate of carbonate-rich<br />

<strong>sediments</strong> <strong>for</strong> trapped charge dating. Radiation Measurements<br />

43, 14–25.<br />

Nichol, S. L. 2002: Morphology, stratigraphy <strong>and</strong> origin of last interglacial<br />

beach ridges at Bream Bay, New Zeal<strong>and</strong>. Journal of<br />

Coastal Research 18, 149–159.<br />

Nichol, S. L., Lian, O. B. & Carter, C. H. 2003: Sheet-gravel evidence<br />

<strong>for</strong> a late Holocene tsunami run-up on beach dunes, Great Barrier<br />

Isl<strong>and</strong>, New Zeal<strong>and</strong>. Sedimentary Geology 155, 129–145.<br />

Nielsen, A., Murray, A. S., Pejrup, M. & Elberling, B. 2006: Optically<br />

stimulated luminescence dating of a Holocene beach ridge plain in<br />

Northern Jutl<strong>and</strong>, Denmark. Quaternary Geochronology 1, 305–312.<br />

Nielsen, K. A., Clemmensen, L. B. & Fornos, J. J. 2004: Middle<br />

Pleistocene magnetostratigraphy: Data from a carbonate aeolian<br />

system, Mallorca, Western Mediterranean. Quaternary Science<br />

Reviews 23, 1733–1756.<br />

Ollerhead, J. & Davidson-Arnott, R. G. D. 1995: The evolution of<br />

Buctouche Spit, New Brunswick, Canada. Marine Geology 124,<br />

215–236.<br />

Ollerhead, J., Huntley, D. J. & Berger, G. W. 1994: <strong>Luminescence</strong><br />

dating of <strong>sediments</strong> from Buctouche Spit, New Brunswick. Canadian<br />

Journal of Earth Sciences 31, 523–531.<br />

Ollerhead, J., Huntley, D. J., Nelson, A. R. & Kelsey, H. M. 2001:<br />

Optical dating of tsunami-laid s<strong>and</strong> from an Oregon <strong>coastal</strong> lake.<br />

Quaternary Science Reviews 20, 1915–1926.<br />

Olley, J. M., De Deckker, P., Roberts, R. G., Fifield, L. K., Yoshida,<br />

H. & Hancock, G. 2004a: Optical dating of deep-sea <strong>sediments</strong><br />

using single grains of quartz: A comparison with radiocarbon. Sedimentary<br />

Geology 169, 175–189.<br />

Olley, J. M., Murray, A. S. & Roberts, R. G. 1996: The effects of<br />

disequilibria in the uranium <strong>and</strong> thorium decay chains on burial<br />

dose rates in fluvial <strong>sediments</strong>. Quaternary Science Reviews 15,<br />

751–760.<br />

Olley, J. M., Pietsch, T. & Roberts, R. G. 2004b: Optical dating of<br />

Holocene <strong>sediments</strong> from a variety of geomorphic setting using<br />

single grains of quartz. Geomorphology 60, 337–358.<br />

Or<strong>for</strong>d, J. D., Murdy, J. M. & Wintle, A. G. 2003: Prograded Holocene<br />

beach ridges with superimposed dunes in north-east Irel<strong>and</strong>:<br />

Mechanisms <strong>and</strong> timescales of fine <strong>and</strong> coarse beach <strong>sediments</strong> decoupling<br />

<strong>and</strong> deposition. Marine Geology 194, 47–64.<br />

Otvos, E. G. 2004: Prospects <strong>for</strong> interregional correlations using<br />

Wisconsin <strong>and</strong> Holocene arid episodes, northern Gulf of Mexico<br />

<strong>coastal</strong> plain. Quaternary Research 61, 105–118.<br />

Otvos, E. G. 2005: Numerical chronology of Pleistocene <strong>coastal</strong> plain<br />

<strong>and</strong> valley development; extensive aggradation during glacial low<br />

sea-levels. Quaternary International 135, 91–113.<br />

Pedoja, K., Dumont, J. F., Lamothe, M., Ortlieb, L., Collot, J. Y.,<br />

Ghaleb, B., Auclair, M., Alvarez, V. & Labrousse, B. 2006a: Plio-<br />

Quaternary uplift of the Manta Peninsula <strong>and</strong> La Plata Isl<strong>and</strong> <strong>and</strong><br />

the subduction of the Carnegie Ridge, central coast Ecuador.<br />

Journal of South American Earth Sciences 22, 1–21.<br />

Pedoja, K., Ortlieb, L., Dumont, J. F., Lamothe, M., Ghaleb, B.,<br />

Auclair, M. & Labrousse, B. 2006b: Quaternary <strong>coastal</strong> uplift along<br />

the Talara Arc (Ecuador, Northern Peru) from new <strong>marine</strong> terrace<br />

data. Marine Geology 228, 73–91.<br />

Peterson, C. D., Stock, E., Price, D. M., Hart, R., Reckendorf, F.,<br />

Erl<strong>and</strong>son, J. M. & Hostetler, S. W. 2007: Ages, distributions, <strong>and</strong><br />

origins of upl<strong>and</strong> <strong>coastal</strong> dune sheets in Oregon, USA. Geomorphology<br />

91, 80–102.<br />

Plater, A. J. & Poolton, N. R. J. 1992: Interpretation of Holocene sea<br />

level tendency <strong>and</strong> intertidal sedimentation in the Tees estuary<br />

using sediment luminescence techniques: A variability study. Sedimentology<br />

39, 1–15.<br />

Porat, N. & Botha, G. A. 2008: The luminescence chronology of dune<br />

development on the Maputal<strong>and</strong> <strong>coastal</strong> plain, southeast Africa.<br />

Quaternary Science Reviews 27, 1024–1046.<br />

Porat, N. & Wintle, A. G. 1995: IRSL dating of aeolianites from the<br />

late Pleistocene kurkar ridge, Israel. INQUA XIV, Berlin, August<br />

1995.<br />

Porat, N., Avital, A., Frechen, M. & Almogi-Labin, A. 2003: Chronology<br />

of upper Quaternary offshore successions from the southeastern<br />

Mediterranean Sea, Israel. Quaternary Science Reviews 22,<br />

1191–1199.<br />

Porat, N., Wintle, A. G. & Ritte, M. 2004: Mode <strong>and</strong> timing of kurkar<br />

<strong>and</strong> hamra <strong>for</strong>mation, central <strong>coastal</strong> plain, Israel. Israel Journal of<br />

Earth Sciences 53, 13–25.<br />

Prescott, J. R. & Fox, P. J. 1990: Dating quartz <strong>sediments</strong> using the<br />

3251C TL peak: New spectral data. Ancient TL 8, 32–34.<br />

Prescott, J. R. & Mojarrabi, B. 1993: Selective bleach: An improved<br />

partial bleach technique <strong>for</strong> finding equivalent doses <strong>for</strong> TL dating<br />

of quartz <strong>sediments</strong>. Ancient TL 11, 27–30.<br />

Prescott, J. R., Huntley, D. J. & Hutton, J. T. 1993: Estimation of<br />

equivalent dose in thermoluminescence dating: The Australian slide<br />

method. Ancient TL 11, 1–5.<br />

Preusser, F., Radies, D., Driehorst, F. & Matter, A. 2005: Late Quaternary<br />

history of the <strong>coastal</strong> Wahiba S<strong>and</strong>s, Sultinate of Oman.<br />

Journal of Quaternary Science 20, 395–405.<br />

Price, D. M., Brooke, B. P. & Woodroffe, C. D. 2001: Thermoluminescence<br />

dating of aeolianites from Lord Howe Isl<strong>and</strong> <strong>and</strong><br />

South-West Western Australia. Quaternary Science Reviews 20,<br />

841–846.<br />

Price, D. M., Bryant, E. A. & Young, R. W. 1999: Thermoluminescence<br />

evidence <strong>for</strong> the deposition of <strong>coastal</strong> <strong>sediments</strong> by<br />

tsunami wave action. Quaternary International 56, 123–128.<br />

Pye, K. & Rhodes, E. G. 1985: Holocene development of an episodic<br />

transgressive dune barrier, Ramsay Bay, North Queeensl<strong>and</strong>, Australia.<br />

Marine Geology 64, 189–202.<br />

Pye, K., Stokes, S. & Neal, A. 1995: Optical dating of aeolian <strong>sediments</strong><br />

from the Sefton coast, northwest Engl<strong>and</strong>. Proceedings of the<br />

Geologists’ Association 106, 281–292.<br />

Rees-Jones, J., Rink, W. J., Norris, R. J. & Litchfield, N. J. 2000:<br />

Optical luminescence dating of uplifted <strong>marine</strong> terraces along the<br />

Akatore Fault near Dunedin, South Isl<strong>and</strong>, New Zeal<strong>and</strong>. New<br />

Zeal<strong>and</strong> Journal of Geology <strong>and</strong> Geophysics 43, 419–424.<br />

Regnauld, H., Mauz, B. & Morzadec-Kerfourn, M. T. 2003: The last<br />

interglacial shoreline in northern Brittany, western France. Marine<br />

Geology 194, 65–77.<br />

Rendell, H. M., Claridge, A. J. & Clarke, M. L. 2007: Late Holocene<br />

Mediterranean <strong>coastal</strong> change along the Tiber delta <strong>and</strong> Roman<br />

occupation of the Laurentine shore, central Italy. Quaternary Geochronology<br />

2, 83–88.


534 Zenobia Jacobs BOREAS<br />

Rhodes, E. J., Singarayer, J. S., Raynal, J.-P., Westaway, K. E. &<br />

Sbihi-Alaoui, F. Z. 2006: New age estimates from the Palaeolithic<br />

assemblages <strong>and</strong> Pleistocene succession of Casablanca, Morocco.<br />

Quaternary Science Reviews 25, 2569–2585.<br />

Richardson, C. A. 2001: Residual luminescence signals in modern<br />

<strong>coastal</strong> <strong>sediments</strong>. Quaternary Science Reviews 20, 887–892.<br />

Rink, W. J. 1998: Quartz luminescence as a light-sensitive indicator of<br />

sediment transport in <strong>coastal</strong> processes. Journal of Coastal Research<br />

15, 148–154.<br />

Rink, W. J. & Forrest, B. 2005: Dating evidence <strong>for</strong> the accretion<br />

history of beach ridges in Cape Canaveral <strong>and</strong> Merritt Isl<strong>and</strong>,<br />

Florida, USA. Journal of Coastal Research 21, 1000–1008.<br />

Rink, W. J. & Pieper, K. D. 2001: Quartz thermoluminescence in a<br />

storm deposit <strong>and</strong> a welded beach ridge. Quaternary Science Reviews<br />

20, 815–820.<br />

Roberts, D. L. & Berger, L. 1997: Last interglacial (c. 117 kyr) human<br />

footprints, South Africa. South African Journal of Science 93,<br />

349–350.<br />

Roberts, D. L., Bateman, M. D., Murray-Wallace, C. V., Carr, A. S.<br />

& Holmes, P. J. 2008: Last Interglacial fossil elephant trackways<br />

dated by OSL/AAR in <strong>coastal</strong> aeolianites, Still Bay, South<br />

Africa. Palaeogeography Palaeoclimatology Palaeoecology 257,<br />

261–279.<br />

Roberts, H. M. & Plater, A. J. 2007: Reconstruction of Holocene<br />

<strong>for</strong>el<strong>and</strong> progradation using optically stimulated luminescence<br />

(OSL) dating: An example from Dungeness, UK. The Holocene 17,<br />

495–505.<br />

Schwebel, D. A. 1984: Quaternary stratigraphy <strong>and</strong> sea-level variation<br />

in the southeast of South Australia. In Thom, B. G. (ed.):<br />

Coastal Geomorphology in Australia, 291–311. Academic Press,<br />

Sydney.<br />

Shaw, A., Holmes, P. J. & Rogers, J. 2001: Depositional l<strong>and</strong><strong>for</strong>ms<br />

<strong>and</strong> environmental change in the headward vicinity of Dias Beach,<br />

Cape Point. South African Journal of Geology 104, 101–114.<br />

Sherwood, J., Barbetti, M., Ditchburn, R., Kimber, R. W. L.,<br />

McCabe, W., Murray-Wallace, C. V., Prescott, J. R. & Whitehead,<br />

N. 1994: A comparative study of Quaternary dating techniques<br />

applied to sedimentary deposits in southwest Victoria, Australia.<br />

Quaternary Science Reviews 13, 95–110.<br />

Shulmeister, J. & Kirk, R. M. 1996: Holocene history <strong>and</strong> a thermoluminescence<br />

based chronology of <strong>coastal</strong> dune ridges near Leithfield,<br />

North Canterbury, New Zeal<strong>and</strong>. New Zeal<strong>and</strong> Journal of<br />

Geology <strong>and</strong> Geophysics 39, 25–32.<br />

Shulmeister, J., Short, S. A., Price, D. M. & Murray, A. S. 1993:<br />

Pedogenic uranium/thorium <strong>and</strong> thermoluminescence <strong>chronologies</strong><br />

<strong>and</strong> evolutionary history of a <strong>coastal</strong> dunefield, Groote Eyl<strong>and</strong>t,<br />

northern Australia. Geomorphology 8, 47–64.<br />

Shulmeister, J., Soons, J. M., Berger, G. W., Harper, M., Holt, S.,<br />

Moar, N. & Carter, J. A. 1999: Environmental <strong>and</strong> sea-level changes<br />

on Banks Peninsula (Canterbury, New Zeal<strong>and</strong>) through three<br />

glaciation – interglaciation cycles. Palaeogeography, Palaeoclimatology,<br />

Palaeoecology 152, 101–127.<br />

Singarayer, J. S. 2002: Linearly Modulated Optically Stimulated <strong>Luminescence</strong><br />

of Sedimentary Quartz: Physical Mechanisms <strong>and</strong> Implications<br />

<strong>for</strong> Dating. D.Phil. thesis, University of Ox<strong>for</strong>d, 340 pp.<br />

Singarayer, J. S. & Bailey, R. M. 2003: Further investigations of the<br />

quartz optically stimulated luminescence components using linear<br />

modulation. Radiation Measurements 37, 451–458.<br />

Singarayer, J. S., Bailey, R. M. & Rhodes, E. J. 2000: Potential of the<br />

slow component of quartz OSL <strong>for</strong> age determination of sedimentary<br />

samples. Radiation Measurements 32, 873–880.<br />

Singhvi, A. K., Deraniyagala, S. U. & Sengupta, D. 1986: Thermoluminescence<br />

dating of Quaternary red-s<strong>and</strong> beds: A case study of<br />

<strong>coastal</strong> dunes in Sri Lanka. Earth <strong>and</strong> Planetary Science Letters 80,<br />

139–144.<br />

Sivan, D. & Porat, N. 2004: Evidence from luminescence <strong>for</strong> Late<br />

Pleistocene <strong>for</strong>mation of calcareous aeolianite (kurkar) <strong>and</strong> palaeosol<br />

(hamra) in the Carmel Coast, Israel. Palaeogeography, Palaeoclimatology,<br />

Palaeoecology 211, 95–106.<br />

Smith, B. W., Prescott, J. R. & Polach, H. 1982: Thermoluminescence<br />

dating of <strong>marine</strong> <strong>sediments</strong> from Spencer Gulf. In Ambrose, W. &<br />

Duerden, P. (eds.): Archaeometry: An Australasian Perspective,<br />

282–289. The Australian National University.<br />

Sommerville, A. A., Hansom, J. D., Housley, R. A. & S<strong>and</strong>erson, D.<br />

C. W. 2007: Optically stimulated luminescence (OSL) dating of<br />

<strong>coastal</strong> aeolian s<strong>and</strong> accumulation in S<strong>and</strong>ay, Orkney Isl<strong>and</strong>s,<br />

Scotl<strong>and</strong>. The Holocene 17, 627–637.<br />

Sommerville, A. A., Hansom, J. D., S<strong>and</strong>erson, D. C. W. & Housley,<br />

R. A. 2003: Optically stimulated luminescence dating of large storm<br />

events in Northern Scotl<strong>and</strong>. Quaternary Science Reviews 22,<br />

1085–1092.<br />

Stokes, S., Ingram, S., Aitken, M. J., Sirocko, F., Anderson, R. &<br />

Leuschner, D. 2003: Alternative <strong>chronologies</strong> <strong>for</strong> Late Quaternary<br />

(Last Interglacial – Holocene) deep sea <strong>sediments</strong> via optical dating<br />

of silt-sized quartz. Quaternary Science Reviews 22, 925–941.<br />

Strickertsson, K., Murray, A. S. & Lykke-Andersen, H. 2001: Optically<br />

stimulated luminescence dates <strong>for</strong> Late Pleistocene <strong>sediments</strong><br />

from Stensnaes, Northern Jutl<strong>and</strong>, Denmark. Quaternary Science<br />

Reviews 20, 755–759.<br />

Sudan, P., Whitmore, G., Uken, R. & Woodborne, S. 2004: Quaternary<br />

evolution of the <strong>coastal</strong> dunes between Lake Hlabane <strong>and</strong><br />

Cape St Lucia, KwaZulu-Natal. South African Journal of Geology<br />

107, 355–376.<br />

Switzer, A. D., Bristow, C. S. & Jones, B. G. 2006: Investigation of<br />

large-scale washover of a small barrier system on the southeast<br />

Australian coast using ground penetrating radar. Sedimentary<br />

Geology 183, 145–156.<br />

Tanaka, K., Hataya, R., Spooner, N. A., Questiaux, D. G., Saito, Y.<br />

& Hashimoto, T. 1997: Dating of <strong>marine</strong> terrace <strong>sediments</strong> by ESR,<br />

TL <strong>and</strong> OSL methods <strong>and</strong> their applications. Quaternary Science<br />

Reviews 16, 257–264.<br />

Tatumi, S. H., Kowata, E. A., Gozzi, G., Kassab, L. R. P., Suguio,<br />

K., Barreto, A. M. F. & Bezerra, F. H. R. 2003: Optical dating of<br />

beachrock, eolic dunes <strong>and</strong> <strong>sediments</strong> applied to sea-level changes<br />

study. Journal of <strong>Luminescence</strong> 102/103, 562–565.<br />

Tejan-Kella, M. S., Chittleborough, D. J., Fitzpatrick, R. W.,<br />

Thompson, C. H., Prescott, J. R. & Hutton, J. T. 1990: Thermoluminescence<br />

dating of <strong>coastal</strong> s<strong>and</strong> dunes at Cooloola <strong>and</strong> North<br />

Stradbrooke Isl<strong>and</strong>, Australia. Australian Journal of Soil Research<br />

28, 465–481.<br />

Thom, B., Hesp, P. & Bryant, E. A. 1994: Last glacial ‘<strong>coastal</strong>’ dunes<br />

in eastern Australia <strong>and</strong> implications <strong>for</strong> l<strong>and</strong>scape stability during<br />

the Last Glacial Maximum. Palaeogeography, Palaeoclimatology,<br />

Palaeoecology 111, 229–248.<br />

Thomas, P. J., Murray, A. S., Granja, H. M. & Jain, M. 2008: Optical<br />

dating of late Quaternary <strong>coastal</strong> deposits in northwestern Portugal.<br />

Journal of Coastal Research 24, 134–144.<br />

Thompson, J., Rink, W. J. & Lopez, G. I. 2007: Optically stimulated<br />

luminescence age of the Old Cedar midden, St. Joseph Peninsula<br />

State Park, Florida. Quaternary Geochronology 2, 350–355.<br />

Tribolo, C., Mercier, N., Selo, M., Valladas, H., Joron, J.-L., Reyss,<br />

J.-L., Henshilwood, C. S., Sealy, J. C. & Yates, R. 2006: TL dating<br />

of burnt lithics from Blombos Cave (South Africa): Further evidence<br />

<strong>for</strong> the antiquity of modern human behaviour. Archaeometry<br />

48, 341–357.<br />

Van Heteren, S., Huntley, D. J., Van de Plassche, O. & Lubberts, R.<br />

K. 2000: Optical dating of dune s<strong>and</strong> <strong>for</strong> the study of sea-level<br />

change. Geology 28, 411–414.<br />

Van Heteren, S., Oost, A. P., Van der Spek, A. J. F. & Elias, E. P. L.<br />

2006: Isl<strong>and</strong>-terminus evolution related to changing ebb-tidal-delta<br />

configuration: Texel, The Netherl<strong>and</strong>. Marine Geology 235, 19–33.<br />

Vogel, J. C., Wintle, A. G. & Woodborne, S. 1999: <strong>Luminescence</strong><br />

dating of <strong>coastal</strong> s<strong>and</strong>s: Overcoming changes in environmental dose<br />

rate. Journal of Archaeological Science 26, 729–733.<br />

Waelbroeck, C., Frank, N., Jouzel, J., Parrenin, F., Masson-<br />

Delmotte, V. & Genty, D. 2008: Transferring radiometric dating of<br />

the last interglacial sea level high st<strong>and</strong> to <strong>marine</strong> <strong>and</strong> ice core<br />

records. Earth <strong>and</strong> Planetary Science Letters 265, 183–194.<br />

Waelbroeck, C., Labeyrie, L., Michel, E., Duplessy, J. C., McManus,<br />

J. F., Lambeck, K., Balbon, E. & Labracherie, M. 2002: Sea-level<br />

<strong>and</strong> deep water temperature changes derived from benthic <strong>for</strong>aminifera<br />

isotopic records. Quaternary Science Reviews 21,<br />

295–305.<br />

Wang, X. L., Lu, Y. C. & Wintle, A. G. 2006: Recuperated OSL dating<br />

of fine-grained quartz in Chinese loess. Quaternary Geochronology<br />

1, 89–100.


BOREAS <strong>Luminescence</strong> <strong>chronologies</strong> <strong>for</strong> <strong>coastal</strong> <strong>and</strong> <strong>marine</strong> <strong>sediments</strong> 535<br />

Wheeler, D. 1995: Character <strong>and</strong> Origin of Anomalous S<strong>and</strong>s, Jervis<br />

Bay, NSW. B.Sc. thesis, University of Wollongong.<br />

Wilson, P., Or<strong>for</strong>d, J., Knight, J., Braley, S. M. & Wintle, A. G. 2001:<br />

Late-Holocene (post-4500 years BP) <strong>coastal</strong> dune development<br />

in Northumberl<strong>and</strong>, northeast Engl<strong>and</strong>. The Holocene 11,<br />

215–229.<br />

Wilson, P., McGourty, J. & Bateman, M. D. 2004: Mid- to late-Holocene<br />

<strong>coastal</strong> dune event stratigraphy <strong>for</strong> the north coast of<br />

Northern Irel<strong>and</strong>. The Holocene 14, 406–416.<br />

Wintle, A. G. & Huntley, D. J. 1979a: Thermoluminescence dating of<br />

deep sea sediment cores. Nature 279, 710–712.<br />

Wintle, A. G. & Huntley, D. J. 1979b: Thermoluminescence dating of<br />

<strong>sediments</strong>. PACT 3, 374–380.<br />

Wintle, A. G. & Huntley, D. J. 1980: Thermoluminescence dating of<br />

ocean <strong>sediments</strong>. Canadian Journal of Earth Sciences 17, 348–360.<br />

Wintle, A. G. & Huntley, D. J. 1982: Thermoluminescence dating of<br />

<strong>sediments</strong>. Quaternary Science Reviews 1, 31–53.<br />

Wintle, A. G., Clarke, M. L., Musson, F. M., Or<strong>for</strong>d, J. & Devoy, R.<br />

J. N. 1998: <strong>Luminescence</strong> dating of recent dunes on Inch Spit,<br />

Dingle Bay, southwest Irel<strong>and</strong>. The Holocene 8, 331–339.<br />

Wood, P. B. 1994: Optically stimulated luminescence dating of a late<br />

Quaternary shoreline deposit, Tunisia. Quaternary Science Reviews<br />

13, 513–516.<br />

Woodroffe, C. D., Bryant, E. A., Price, D. M. & Short, S. A. 1992:<br />

Quaternary inheritance of <strong>coastal</strong> l<strong>and</strong><strong>for</strong>ms, Coburg Peninsula,<br />

Northern Territory. Australian Geographer 23, 101–115.<br />

Woodroffe, C. D., Kennedy, D. M., Brooke, B. P. & Dickson, M. E.<br />

2006: Geomorphological evolution of Lord Howe Isl<strong>and</strong> <strong>and</strong> carbonate<br />

production at the latitudinal limit to reef growth. Journal of<br />

Coastal Research 22, 188–201.<br />

Woodroffe, C. D., Murray-Wallace, C. V., Bryant, E. A., Brooke, B.<br />

P., Heijnis, H. & Price, D. M. 1995: Late Quaternary sea-level<br />

highst<strong>and</strong>s in the Tasmanian Sea: Evidence from Lord Howe<br />

Isl<strong>and</strong>. Marine Geology 125, 61–72.<br />

Wu, Z., Wang, W., Tan, H. H. & Xu, F. Y. 2000: The age of the ‘old<br />

red s<strong>and</strong>’ on the coasts of south Fujian <strong>and</strong> west Guangdong,<br />

China. Chinese Science Bulletin 45, 1216–1221.<br />

Yim, W.-S., Hilgers, A., Huang, G. & Radtke, U. 2008: Stratigraphy<br />

<strong>and</strong> optically stimulated luminescence dating of subaerially<br />

exposed Quaternary deposits from two shallow bays in Hong<br />

Kong, China. Quaternary International, doi: 10.1016/j.quaint.2007.<br />

07.004.<br />

Yim, W.-S., Price, D. M. & Choy, A. M. S. F. 2002: Distribution of<br />

moisture contents <strong>and</strong> thermoluminescence ages in an inner shelf<br />

borehole from the new Hong Kong International airport site,<br />

China. Quaternary International 92, 35–43.<br />

Yoshida, H., Roberts, R. G. & Olley, J. M. 2003: Progress towards<br />

single-grain optical dating of fossil mud-wasp nests <strong>and</strong> associated<br />

rock art in northern Australia. Quaternary Science Reviews 22,<br />

1273–1278.<br />

Yoshida, H., Roberts, R. G., Olley, J. M., Laslett, G. M. & Galbraith,<br />

R. F. 2000: Extending the age range of optical dating using single<br />

‘supergrains’ of quartz. Radiation Measurements 32, 439–446.<br />

Young, R. W., Bryant, E. A. & Price, D. M. 1993a: Last interglacial<br />

sea levels on the south coast of New South Wales. Australian Geographer<br />

24, 72–75.<br />

Young, R. W., Bryant, E. A. & Price, D. M. 1995: The imprint of<br />

tsunami in Quaternary <strong>coastal</strong> <strong>sediments</strong> of southeastern Australia.<br />

Bulgarian Geophysical Journal 21, 24–32.<br />

Young, R. W., Bryant, E. A. & Price, D. M. 1996: Catastrophic wave<br />

(tsunami?) transport of boulders in southern New South Wales,<br />

Australia. Zeitschrift für Geomorphologie 40, 191–207.<br />

Young, R. W., Bryant, E. A., Price, D. M., Dilek, S. Y. & Wheeler, D.<br />

J. 1997: Chronology of Holocene tsunamis on the southeastern<br />

coast of Australia. Transactions of the Japanese Geomorphological<br />

Union 18, 1–19.<br />

Young, R. W., Bryant, E. A., Price, D. M., Wirth, L. M. & Pease, M.<br />

1993b: Theoretical constraints <strong>and</strong> chronological evidence of Holocene<br />

<strong>coastal</strong> development in central <strong>and</strong> southern New South<br />

Wales. Geomorphology 7, 317–329.<br />

Z<strong>and</strong>er, A., Degering, D., Preusser, F., Kasper, H. U. & Bruckner, ¨ H.<br />

2007: Optically stimulated luminescence dating of sublittoral <strong>and</strong><br />

intertidal <strong>sediments</strong> from Dubai, UAE: Radioactive disequilibria in<br />

the uranium decay series. Quaternary Geochronology 2, 123–128.<br />

Zazo, C., Mercier, N., Silva, P. G., Dabrio, C. J., Goy, J. L., Roquero,<br />

E., Soler, V., Borja, F., Lario, J., Polo, D. & de Luque, L.<br />

2005: L<strong>and</strong>scape evolution <strong>and</strong> geodynamic controls in the Gulf of<br />

Cadiz (Huelva coast, SW Spain) during the late Quaternary. Geomorphology<br />

68, 269–290.<br />

Zhang, J.-F., Fan, C.-F., Wang, H. & Zhou, L.-P. 2007: Chronology<br />

of an oyster reef on the coast of Bohai Bay, China: Constraints<br />

from optical dating using different luminescence signals from fine<br />

quartz <strong>and</strong> polymineral fine grains of <strong>coastal</strong> <strong>sediments</strong>. Quaternary<br />

Geochronology 2, 71–76.<br />

Zhang, J.-F., Yuan, B. Y. & Zhou, L.-P. 2008: <strong>Luminescence</strong> chronology<br />

of ‘Old Red S<strong>and</strong>’ in Jinjiang <strong>and</strong> its implications <strong>for</strong> optical<br />

dating of <strong>sediments</strong> in South China. Chinese Science Bulletin 53,<br />

591–601.<br />

Zhou, L.-P., Williams, M. A. J. & Peterson, J. A. 1994: Late<br />

Quaternary aeolianites, palaeosols <strong>and</strong> depositional environments<br />

on the Nepean Peninsula, Victoria, Australia. Quaternary Science<br />

Reviews 13, 225–239.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!