Towards a science of past disasters

28
ORIGINAL PAPER Towards a science of past disasters Felix Riede Received: 13 June 2013 / Accepted: 25 October 2013 / Published online: 6 November 2013 Ó Springer Science+Business Media Dordrecht 2013 Abstract It is widely recognised that natural disasters emerge in the interplay between extreme geophysical events and the human communities affected by them. Whilst detailed natural scientific knowledge of a given event is critical in understanding its impacts, an equally thorough understanding of the affected communities, their economies, ecologies, religious structures, and how all of these have developed over time is arguably as important. Many extreme events leave methodologically convenient traces in the geo- logical and archaeological records in the form of discrete stratigraphic layers often asso- ciated with both accurate and precise dates. This paper focuses on volcanic eruptions and draws on matched case studies to illustrate the usefulness of a two-step, quasi case–control comparative method for examining vulnerability and impacts in the near- and far-fields of these eruptions. Although issues of data resolution often plague the study of past disasters, these limitations are counterbalanced by the access to unique long-term information on societies and their material expressions of livelihood, as well as a similarly long-term perspective on the critical magnitude/frequency relationship of the geophysical trig- ger(s) in question. By drawing together aspects of contemporary Disaster Risk Reduction research, archaeology, and volcanology, this paper sketches out a methodological roadmap for a science of past disasters that aims to be relevant for not only understanding vul- nerabilities and impacts in the deep past, but for also better understanding vulnerability in the present. Keywords Archaeology Past disaster science Natural experiments of history Laacher See eruption Thera eruption Volca ´n Ilopango Eyjafjallajo ¨kull F. Riede (&) Laboratory for Past Disaster Science (LaPaDiS), Department of Culture and Society (Materials, Culture and Heritage), Aarhus University, Campus Moesga ˚rd, 8270 Højbjerg, Denmark e-mail: [email protected] 123 Nat Hazards (2014) 71:335–362 DOI 10.1007/s11069-013-0913-6

Transcript of Towards a science of past disasters

ORI GIN AL PA PER

Towards a science of past disasters

Felix Riede

Received: 13 June 2013 / Accepted: 25 October 2013 / Published online: 6 November 2013� Springer Science+Business Media Dordrecht 2013

Abstract It is widely recognised that natural disasters emerge in the interplay between

extreme geophysical events and the human communities affected by them. Whilst detailed

natural scientific knowledge of a given event is critical in understanding its impacts, an

equally thorough understanding of the affected communities, their economies, ecologies,

religious structures, and how all of these have developed over time is arguably as

important. Many extreme events leave methodologically convenient traces in the geo-

logical and archaeological records in the form of discrete stratigraphic layers often asso-

ciated with both accurate and precise dates. This paper focuses on volcanic eruptions and

draws on matched case studies to illustrate the usefulness of a two-step, quasi case–control

comparative method for examining vulnerability and impacts in the near- and far-fields of

these eruptions. Although issues of data resolution often plague the study of past disasters,

these limitations are counterbalanced by the access to unique long-term information on

societies and their material expressions of livelihood, as well as a similarly long-term

perspective on the critical magnitude/frequency relationship of the geophysical trig-

ger(s) in question. By drawing together aspects of contemporary Disaster Risk Reduction

research, archaeology, and volcanology, this paper sketches out a methodological roadmap

for a science of past disasters that aims to be relevant for not only understanding vul-

nerabilities and impacts in the deep past, but for also better understanding vulnerability in

the present.

Keywords Archaeology � Past disaster science � Natural experiments

of history � Laacher See eruption � Thera eruption � Volcan Ilopango �Eyjafjallajokull

F. Riede (&)Laboratory for Past Disaster Science (LaPaDiS), Department of Culture and Society (Materials,Culture and Heritage), Aarhus University, Campus Moesgard, 8270 Højbjerg, Denmarke-mail: [email protected]

123

Nat Hazards (2014) 71:335–362DOI 10.1007/s11069-013-0913-6

1 Introduction

It is widely recognised that natural disasters are the result of the complex interaction

between extreme geophysical events and the human communities affected by them (Hewitt

1983; Oliver-Smith 1999; Quarantelli 1995; Felgentreff and Glade 2008). Whilst detailed

natural scientific knowledge of a given event is critical in understanding its impacts, an

equally thorough understanding of the affected communities, their economies, ecologies,

religious structures, and how all of these have developed over time can be said to be as

important—especially if the aim is to not only retrospectively relate post-event impacts to

pre-event patterns of vulnerability, but to use such analyses to reflect on (a) the societal

impacts of extreme events and (b) how efficiently and effectively to prepare for future

calamities. Arguably, history matters when investigating the relationship between human

communities and extreme events (Bankoff 2004). Following Garcıa-Acosta (2002: 65)

‘‘emphasis should be placed on understanding the surrounding and prior sociocultural

context and vulnerability to the effects of a certain hazard. Examining one of the key

theoretical issues in any disaster research—the multidimensionality of disasters as

expressed in the concept of socially constructed vulnerability—deepens our knowledge of

hazards themselves: to determine the cause of calamitous incidence, recurrence, and

probability; to differentiate scale, intensity, and duration; and to understand how to face

disasters or avoid them’’.

Since the ‘‘radical critique’’ of the 1980s (O’Keefe et al. 1976; Hewitt 1983) similar

rallying calls for routinely including historical dimensions in the study of vulnerability

have been sounded repeatedly and from different disciplinary corners (Grayson and Sheets

1979; Leroy 2006; Bankoff 2004), yet they have only sporadically been answered (see

chapters in Cooper and Sheets 2012; Mauch and Pfister 2009; Janku et al. 2012; Schenk

2009). This continuing emphasis on technocratic risk reduction solutions finds one of its

perhaps most stark reflections in the allocation of research funds: For example, despite

there being a fair share of social science subjects amongst the *30 stakeholder disciplines

in the field of Disaster Risk Reduction (Alexander 1997), approximately 95 % of the

available funding goes towards natural science and engineering projects (Alexander 1995).

Arguably, the measures advocated by the latter disciplines have not, however, resulted in

the hoped-for disaster panacea (Oliver-Smith 2004), and cost-effective and lasting vul-

nerability reduction and disaster mitigation—especially in contexts where technological

relief measures are prohibitively expensive to implement or where the nature of the

calamity makes such implementations difficult or impossible (e.g. large volcanic eruptions;

see Schmincke and Hinzen 2008)—can be greatly improved via a more balanced approach

that integrates both natural and social sciences.

Economic development and its attendant parameters such as wealth inequality, market

distribution structures, population increase/change, and urbanisation are often put forward

as some of the key variables creating and amplifying vulnerability (e.g. Cutter et al. 2003;

Birkmann et al. 2013). Interestingly, archaeological and historical analyses of disaster

impacts diverge in their general conclusions with regard to the way in which development

interacts with vulnerability and resilience: Whilst historians tend to argue that development

ultimately increases resilience and indeed that disasters spur on development (Pfister 2009;

Helbling 2006), archaeologists see the emergence of centralised political systems and

market economies as a key factor that leads to increased vulnerability (Fitzhugh 2012;

Sheets 1999, 2001, 2008, 2012). Whilst seemingly in opposition, these diverging obser-

vations can to some degree be reconciled with reference to Gilbert White (1974) who

already noted a long time ago that effective post-industrial or comprehensive responses to

336 Nat Hazards (2014) 71:335–362

123

disasters incorporate elements of pre-industrial as well as industrial social configurations

(Table 1). Yet the many known cases where, in particular, colonial interventions have led

to increased vulnerability amongst indigenous populations suggest, however, that not all

forms of development benefit resilience, whereas—vice versa—not all forms of indigenous

or pre-industrial response can be seen as effective or useful (Chester et al. 2012). In

addition, White’s tripartite distinction of social responses into pre-industrial, industrial, and

post-industrial does not easily facilitate a more detailed assessment of vulnerabilities

amongst traditional societies, which clearly differ along many more salient dimensions

than merely economic organisation. Given the considerable recent interest in the potential

of traditional knowledge informing participatory approaches to risk reduction and coping

in general (Wisner 2010; Adger 2006; Lorenz 2013) and also with specific reference to

volcanic hazards (De Belizal et al. 2012; Cronin et al. 2004b; Cronin et al. 2004a; Cashman

and Cronin 2008), deciding which pre-industrial and industrial elements are best incor-

porated into a comprehensive disaster response approach remains a pressing empirical

question.

This paper introduces and discusses a semi-formalised method for how historical and

archaeological source data can be used to explore social, ecological, and place vulnera-

bility as well as resilience. Disasters have often been likened to social laboratories (Garcıa-

Acosta 2002; Oliver-Smith 1996; Grayson and Sheets 1979), and this notion is here

developed further by explicitly linking it to formal aspects of the case–control study

design. The application of such comparative methods in disciplines concerned with the past

has been dubbed ‘‘natural experiments of history’’ (Diamond and Robinson 2010b).

Importantly, many extreme events (e.g. earthquakes, tsunamis/storm surges, volcanic

eruptions) leave methodologically convenient stratigraphic traces in the geological and

archaeological records that facilitate the precise and accurate correlation between sites and

Table 1 Pre-industrial and industrial responses to natural hazards. After White (1974) and Chester et al.(2012)

Pre-industrial Industrial

Response types and characteristics

Adjustment range Wide Restricted

Actors Individuals, households, smallgroups, communities

Authorities, authority-coordinatedgroups

Relation to nature Harmonisation with Technological control over

Capital investment Low High

Spatial variability inresponses

High Low

Response flexibility High Low

Loss perception Perceived as inevitable Losses may/should be reduced bygovernment action, technology,science and development

Time-depth Deep; where there is previoushazard knowledge, traditionalresponses are evolved

Shallow; industrial responses firstemerge from mid-nineteenth centuryonwards and often suppress or replace(especially in marginal and colonisedareas) traditional responses

Post-industrial or comprehensive response approaches combine the most effective elements of both,although, arguably, such mitigation measures have not been implemented anywhere in the world yet

Nat Hazards (2014) 71:335–362 337

123

regions. This temporal control allows near- and far-field effects to be investigated, and

climatic and environmental records to be directly linked to sequences of culture change.

In a first illustration of such an approach, this paper focuses on volcanic eruptions and

examines one particular case study in detail—the Laacher See eruption (Germany) in the

very late twelfth or early eleventh millennium BCE (Before Common Era)—to showcase

the usefulness of a case–control comparative method for examining geographically dif-

ferentiated vulnerabilities and impacts in the eruption’s near- and far-fields. This analysis

proceeds in a two-step manner by first investigating the eruption’s cultural effects

synchronically across multiple regions and thereafter by placing these effects into a

broader diachronic perspective. This second analytical step relaxes the requirements of

comparison but in doing so builds a bridge between the archaeological case study and the

present day by pointing to other salient case studies of how volcanic eruptions at different

times in Europe’s deep and recent past—notably the eruption of Thera around 1610 BCE,

the large eruption of Volcan Ilopango (El Salvador) in the sixth century CE, and the

somewhat smaller Eyjafjallajokull eruption of 2010 CE—have impacted societies at dif-

ferent levels of sociopolitical complexity. In combination, the results of this two-step

analysis strongly implicate factors other than the direct fallout of volcanic ash as critical in

causing post-event demographic and culture change. In turn, it is these traits (population

density, network position, resource diversity) that arguably structure pre-event

vulnerability.

The case studies employed here collectively underline that calamities such as volcanic

eruptions can lead to cascading effects that reverberate through social and demographic

networks at variable speeds. Such events, it seems, can have long-term social and political

legacies, and their effects are often indirect, mediated by culturally specific components

such as religion, and that these effects can occur or indeed be amplified in the far-field.

Although issues of data resolution often plague the study of past disasters, these limitations

are arguably counterbalanced by access to unique long-term information on societies and

their material expressions of livelihood, as well as a similarly long-term perspective on the

critical magnitude/frequency relationship of the natural hazards in question. Finally, it is

argued that information from past calamities may be used to inform planning for future

extreme events. Clarke (1999, 2006, 2008a, b), for instance, has long argued that not only

probable but also possible events should be the subject of serious debate and planning

efforts. Whilst such ‘‘possibilism’’ carries with it the danger of hysteria, archaeological and

historical data can be effectively used to ‘‘discipline possibilistic reasoning’’ (Clarke 2007:

192) by offering historically informed, evidence-based information on both the geophys-

ical as well as sociocultural parameters of past extreme events.

2 Materials and methods

2.1 The Laacher See eruption

In late spring or early summer approximately 13,000 years ago, the Laacher See volcano

erupted cataclysmically. The Laacher See volcano is part of the East Eifel volcanic field,

Rhenish Shield (Germany), and the its caldera is located in the now densely settled

Neuwied Basin between the cities of Bonn and Koblenz (Schmincke 2006; Schmincke

et al. 1999). With a calculated eruption magnitude of M = 6.2 and an eruption intensity

I C 11.5 (see Pyle 2000), the Laacher See eruption (LSE) ranks globally as a very intense

volcanic event of moderate-to-large proportion. The rising magma’s interaction with

338 Nat Hazards (2014) 71:335–362

123

groundwater resulted in a highly explosive phreatomagmatic eruption that was at its most

intense over a period of a few days or weeks, but probably lasted several months in all (Litt

et al. 2008). During the eruption, a strongly zoned phonolitic magma chamber was tapped

resulting in a series of ejecta that are petrologically and volcanologically well described

(van den Bogaard and Schmincke 1985, 1984; Worner and Schmincke 1984a, b). The

eruption column varied in height over time and has been estimated to have reach-

ed B 40 km (van den Bogaard et al. 1990). Climate models suggest that the atmospheric

input of aerosols from this eruption resulted in altered Northern Hemisphere weather

patterns for some years (Graf and Timmreck 2001; Textor et al. 2003), which is duly

reflected in terrestrial and lacustrine records across Europe (Birks and Lotter 1994; Merkt

and Muller 1999; de Klerk et al. 2008).

An estimated near-field area of[1,400 km2 was completely covered in pyroclastic flow

and fallout deposits ranging from 50 to 1 m thickness. At the nearby River Rhine, these

deposits built up to form a dam, which in turn led to the formation of a substantial lake and

widespread upstream flooding and attendant downstream channel drying. Nearest to the

eruptive centre, the River Rhine today has an average discharge rate of *2,000 m3/s

(Kwadijk 1991). Assuming a roughly similar discharge rate for the period around the end

of the last ice age, the respective drying and flooding caused by the LSE dam would have

had dramatic and highly visible consequences for regions both up- and perhaps especially

downstream of the Eifel. Water-rafted pumice and other overbank features suggest that this

dam likely broke during or shortly after the eruption, causing one or several gargantuan

laharlike flood waves (Park and Schmincke 1997, 2009), which possibly led to a major

reconfiguration of the Rhine delta (Janssens et al. 2012; Erkens et al. 2011). Depending on

the height of the eruption column and prevailing winds, volcanic ash (tephra) from this

eruption was transported over large parts of Europe, from Italy in the south to the margins

of the inland ice near present day Gotland in the north and from the Ardennes in the west to

northern central Poland in the east (Fig. 1). Upwards of 600 data points for the occurrence

of Laacher See tephra across Europe are currently known. Tables 2 and 3 summarise key

eruption parameters.

2.2 Affected communities

Archaeologists working in the deep past are regularly confronted with a highly fragmented

record of prehistoric peoples’ lives (see, for example, Scarre 2005). The most common

remains are tools fashioned out of stone or durable organic raw materials. The shape and

technological details of these tools together with their distribution in space and time are

used to define archaeological ‘‘cultures’’. These entities most certainly are not identical to

the kinds of cultures defined by sociologists or anthropologists in the present day, but they

can be thought of as communities of knowledge and know-how shared largely, but by no

means exclusively, within extended family groups (Riede 2011a). These groupings almost

certainly also had languages and other less durable cultural features in common; material

culture is treated as an effective proxy of immaterial commonalities (see, for instance,

papers in Roberts and Vander Linden 2011).

Europe at the end of the last ice age was occupied at low population densities by groups

of nomadic hunter-gatherers belonging to the cultural tradition of the so-called Upper

Magdalenian culture, which in northern Europe is known by the term Federmessergruppen

(= Penknife groups), after their characteristic arrowhead design similar to knives used to

sharpen feather quill pens (Fig. 2a). The Magdalenian as such is associated with the human

re-colonisation of northern Europe at the end of the last ice age (Gamble et al. 2004).

Nat Hazards (2014) 71:335–362 339

123

Fig. 1 The currently known distribution of tephra and other volcanic products from the Laacher Seeeruption superimposed on the approximate palaeogeography of Europe at the end of the last ice age. Thelocation of the Laacher See edifice is marked by a larger circle, small circles denote locations with knownairfall tephra, triangles locations with fluvially transported ejecta. Three distal find localities in northernItaly are not shown

Table 2 The number of LaacherSee tephra (LST) occurrencescurrently known from Europe,updated from Riede et al. (2011)and Riede and Thastrup (2013)

Note the large number ofcountries directly affected andtheir central position within thepolitical and economicgeography of Europe

No. Country NLST

1 Austria 2

2 Belgium 9

3 Denmark 2

4 France 100

5 Germany 439

6 Italy 2

7 Luxembourg 13

8 Netherlands 6

9 Poland 4

10 Sweden 2

11 Switzerland 34

SUM 613

340 Nat Hazards (2014) 71:335–362

123

Different regions have had their own settlement trajectories since initial colonisation, and

there is some regional variation in resource diversity and use (e.g. Jochim et al. 1999).

Overall, however, the Upper Magdalenian/Federmessergruppen cultural complex is seen as

Table 3 Summary of information on the Laacher See eruption and its climatic, environmental, and culturalimpact

Volcanic zone East Eifel volcanic field, RhenishShield, western Germany

Date estimates 10.970 calendar years BCE10.950 ± 560 (40Ar/39Ar) BCE10.930 ± 40 varve years BCE11.080 calendar years BCE

Baales et al. (2002)van den Bogaard (1995)Brauer et al. (1999)van Raden et al. (in press)

Correlated geophysical,cosmogenic, and climaticeffects

Acid rain, increased rain fall,unseasonal lightning strikes,reduction in solar radiation,drop in temperature, pressurewaves/earthquakes

van den Bogaard et al.(1990), Graf and Timmreck(2001), Schmincke (2006),Meischner and Gruger(2008)

Wind direction and falloutlobes in order of volume

NE[S[SW van den Bogaard andSchmincke (1984, 1985)

Maximum height of ashcolumn

40 km van den Bogaard andSchmincke (1985)

Minimum height of ashcolumn

20 km Schmincke et al. (1999)

Volume of extruded magma 20.0 km3 (6.3 km3 dense rockequivalent)

Schmincke et al. (1999)

Magnitude (M) 6.2 After Mason et al. (2004) andPyle (2000)

Intensity (I) C11.5

Volcanic Explosivity Index 5–6 After Newhall and Self(1982)

Discharge rate estimates 3–5 9 108 kg/s Schmincke (2006)

Ejecta temperatures 800–880 �C (magma)250 �C (pyroclastic flows)

Worner and Schmincke(1984b). Schmincke (2006)

Sulphur injected into theatmosphere

1 9 1014 g SO2

2 9 1013 g SO2

2–15 9 1012 g S

van den Bogaard et al. (1990)Harms and Schmincke (2000)Schmincke et al. (1999)

Minimum area covered bypyroclastic currents

[1,400 km2 van den Bogaard andSchmincke (1984)

Minimum area affected bytephra fallout

700,000 km2 (C1 mm thickness)225,000 km2 (C5 mm thickness)[230,000–325,000 km2 (variable

thickness)

van den Bogaard andSchmincke (1985)

Fisher and Schmincke (1984)Riede et al. (2011)

Northern Hemisphere cooling 0.5 �C1–2 �C

van den Bogaard et al. (1990)Graf and Timmreck (2001)

High-latitude ([60�N)amplifying factor forcooling

?4 (late winter)/-4 (latesummer) �C

Graf and Timmreck (2001)

Effects on contemporaneoushunter-gatherercommunities

Parts of the North European Plainunder tephra fallout avoided

Riede (2008, 2007)

BCE before common era

Nat Hazards (2014) 71:335–362 341

123

remarkable homogeneous: ‘‘a major degree of uniformity appears to characterise these

industries, covering a very large area’’ (De Bie and Vermeersch 1998: 37).

The presence of adornments and other non-utilitarian objects suggests that not only

mundane domestic material culture, but also symbols that likely functioned as identify

markers were widely shared in northern Europe at the time (e.g. Alvarez-Fernandez 2009),

although these same items may also have played a key role in emerging social differen-

tiation and political manoeuvring (Schwendler 2012). In addition, the relative uniformity

of the material found at most sites and the nature of the dwellings discovered there

suggest—in analogy with the ethnographic record—the presence of extended family

groups of mixed age and gender that formed small and largely self-sufficient domestic

units (Gelhausen et al. 2004; Loew 2009). Population densities (Bocquet-Appel et al. 2005)

and resource use (Eriksen 1996) varied somewhat from region to region, but these subtly

different regional manifestations effectively all constitute variations on a ‘‘background

theme linking the whole area of investigation’’ (Weniger 1989: 365).

This homogeneity decreases markedly after about 11,000 BCE. In southern Scandina-

via, for instance, the archaeological record reveals the emergence of a geographically

tightly circumscribed community, the so-called Bromme culture, named after the epony-

mous excavation site where it was first recognised (Mathiassen 1946). This archaeological

culture is characterised by (1) an impoverished tool repertoire consisting of (2)

Fig. 2 a Key examples of material culture associated with the Federmessergruppen and its approximategeographic distribution: an elk figurine of amber (e.g. Veil et al. 2012) remains of fishing hooks (e.g. Pasda2001), harpoons (e.g. Baales 2002), sandstone abraders used to fashion arrow and spear shafts (e.g. De Bieand Caspar 2000), and a range of chipped stone tools used as projectile tips for both javelins and arrows andfor working hides, antler and bone. b Key examples of material culture associated with the Bromme cultureand its approximate geographic distribution: three types of chipped stone tool used as projectile tips for onlyjavelins and for working hides, antler and bone, sandstone abraders used to fashion spear shafts (Riede2012a). Objects not to scale. The location of the Laacher See volcano is noted on both maps

342 Nat Hazards (2014) 71:335–362

123

predominantly chunky tools as well as (3) by its spatially limited occurrence (Fig. 2b).

These three traits go strongly against the grain of contemporaneous cultural development

and thus are in need of explanation. Attempts to correlate the emergence of this culture

with general patterns of climate or environmental change or with the abundance of high-

quality flint resources in the area (Sørensen 2010; Brinch Petersen 2009) fail to account

satisfactorily for its specific characteristics and for the wider implications of the evident

loss of a key technological feature of the time, bow and arrow technology (Riede 2009;

Dev and Riede 2012).

It has been noted that there is a close spatio-temporal correlation of the Bromme culture

with the tephra fallout from the Laacher See eruption. With respect to geography, southern

Scandinavia is framed by the tephra fallout, and the Bromme culture appears to avoid

affected areas (Riede 2007, 2008). With respect to chronology, the Bromme culture dates

to after this eruptive event (Fischer et al. 2013; Riede and Edinborough 2012). In analogy

with ethnographic examples, it has been argued that the LSE led to demographic fluctu-

ations and disruptions of contemporaneous travel and communication routes and therefore

a fragmentation of the settlement area in the period after the eruption, eventually leading to

the observed culture change (Riede 2007, 2008, 2012b).

In other areas, the period after 11,000 BCE is characterised by either more subtle

changes or no changes at all. South-western Germany/northern Switzerland sees a slight

variation in the material culture summarised under the label ‘‘Fursteiner group’’ (Bandi

1968; Nielsen 2009). Regions such as the British Isles, the Belgian uplands, and north-

central Germany as well as the Rhineland appear to become largely or entirely depopulated

for some time. From these areas, several stratigraphic sequences preserving both archae-

ological layers and layers of LST directly above are known (Riede and Thastrup 2013). In

contrast, the cultural sequences from the Netherlands and from the Paris Basin region are

unbroken throughout this period and imply gradual changes as well as perhaps even

demographic increase (Riede 2008). Whilst only the southern Scandinavian record has

been systematically investigated in relation to a possible impact of the Laacher See

eruption, the patterns observed in the remaining regions do allow an initial and exploratory

comparison. Figure 3 and Table 4 schematically summarise the cultural sequences for

these regions, places them in relation to a general temperature proxy (isotopic variations in

Fig. 3 Schematic of the matched case–control studies of the impact of the Laacher See eruption in differentregions. *indicates that the southern part of Benelux (Belgium, Luxembourg) may have been abandoned,whereas the northern part (The Netherlands) may have received a population influx at this time. Of particularrelevance here is the period between 12.000 and 10.000 BCE. The temperature proxy curve and phasenotation follows Lowe et al. (2008). GI = Greenland Interstadial (warm), GS = Greenland Stadial (cold).The cooler GI-1b corresponds to the Intra-Allerød Cold Phase (see Fig. 6 below)

Nat Hazards (2014) 71:335–362 343

123

Ta

ble

4S

um

mar

yo

fth

ev

aria

ble

su

sed

toin

ves

tig

ate

loca

tion

alan

dso

cio

-eco

log

ical

vu

lner

abil

ity

atth

een

do

fth

ela

stic

eag

e

Imp

acta

Tep

hra

rece

ived

bL

atit

ud

eD

ista

nce

from

ven

tR

eso

urc

ed

iver

sity

Tim

esi

nce

colo

nis

atio

nN

etw

ork

po

siti

on

Po

pu

lati

on

den

sity

Sel

ecte

dk

eyre

fere

nce

s

SW

Ger

man

yan

dN

Sw

itze

rlan

dP

has

ech

ang

e(1

)D

ista

l(1

)4

7.5

Mo

der

ate

(1)

Hig

h(1

)0

Cen

tral

(0)

Hig

h(2

)N

iels

en(2

00

9),

Lee

sch

etal

.(2

01

2)

Rhin

elan

dD

epopula

tion

(3)

Pro

xim

al(3

)50.0

Clo

se-b

y(0

)H

igh

(1)

1C

entr

al(0

)H

igh

(2)

Baa

les

etal

.(2

00

2),

Str

eet

etal

.(2

01

2)

Par

isB

asin

No

ne

(0)

No

ne

(0)

49

.0M

oder

ate

(1)

Hig

h(1

)2

Cen

tral

(0)

Hig

h(2

)B

od

u(1

99

8),

Deb

ou

tet

al.

(20

12),

Co

ud

ret

and

Fag

nar

t(1

99

7)

NC

Ger

man

yD

epo

pula

tio

n(3

)M

edia

l(2

)5

2.0

Mo

der

ate

(1)

Lo

w(0

)3

Per

iph

eral

(1)

Lo

w(1

)K

uss

ner

(20

10),

Gru

nb

erg

(20

06),

Rie

de

(20

12

b)

Ben

elu

xN

on

e(0

)N

on

e(0

)5

2.0

Mo

der

ate

(1)

Lo

w(0

)3

Per

iph

eral

(1)

Hig

h(2

)D

eB

iean

dV

erm

eers

ch(1

99

8),

Sta

per

t(2

00

0)

Bri

tish

Isle

sD

epo

pula

tio

n(3

)N

on

e(0

)5

1.0

Far

-fiel

d(2

)H

igh

(1)

4P

erip

her

al(1

)L

ow

(1)

Pet

titt

(20

08),

Pet

titt

and

Wh

ite

(20

12)

SS

can

din

avia

Cu

ltu

rech

ang

e(2

)N

on

e(0

)5

5.0

Far

-fiel

d(2

)L

ow

(0)

5P

erip

her

al(1

)L

ow

(1)

Rie

de

(20

07,

20

08)

aM

ark

sth

eo

utc

om

ev

aria

ble

,bm

ark

sth

ep

ertu

rbat

ion

var

iab

le.

Tim

esi

nce

colo

nis

atio

nis

sim

ply

ord

ered

dep

endin

go

nth

eti

min

go

fth

efi

rst

hu

man

re-c

olo

nis

atio

nin

each

reg

ion

.T

he

nu

mb

erg

iven

inb

rack

ets

afte

rea

chv

aria

ble

corr

esp

on

ds

toth

eco

din

gu

sed

for

sub

sequ

ent

anal

yse

s

344 Nat Hazards (2014) 71:335–362

123

ice-trapped oxygen) from the Greenland ice cores, and shows which of these were affected

by fallout tephra from the Laacher See eruption.

2.3 Natural experiments of history, the comparative method, and case–control studies

Any discipline concerned with historical data (e.g. evolutionary biology, palaeontology,

epidemiology, geology, historical linguistics, economy, political science, astronomy,

anthropology, and archaeology) cannot for practical and/or ethical reasons conduct labora-

tory experiments (see Galavotti 2003; Dunning 2012; Morgan 2013). Instead, these disci-

plines resort to careful description, analysis, and comparison, especially when attempting to

generalise across particular analyses and when inferring causality. They pragmatically

employ what has been termed ‘‘natural experiments of history’’ (Diamond and Robinson

2010b). In evolutionary biology, for instance, the (phylogenetic) comparative method has

long been formalised and implemented in computational applications (Harvey and Pagel

1991). It constitutes ‘‘one of biology’s most enduring sets of techniques for investigating

evolution and adaptation’’ (Pagel and Meade 2005: 235). The same comparative methods are

also occasionally employed by anthropologists interested in inferring, for instance, cultural

changes on the basis of contemporaneous behaviour traits and their distribution (see Mace and

Holden 2005) or by archaeologists interested in sequences of coupled technological changes

(e.g. Riede 2011b). When used in the social sciences, the method is described as ‘‘the retro-

fitting by social scientists of events that have happened in the social world into the traditional

forms of field or randomised trial experiments’’ (Morgan 2013:341). Although an appreci-

ation of the potential inferential power of this method is not universal, its use is becoming

more widespread (Dunning 2008), and comparative, natural experimental studies are being

conducted with increasing frequency in a range of social science disciplines. However, such a

methodology has not yet been applied to studies of natural hazard impacts.

In epidemiology, this kind of comparative methodology has a long history and is known

as the case–control study design. It is one of the discipline’s major tools for inferring

causality retrospectively. In contrast to the comparative method in biology and anthro-

pology, however, epidemiologists are not concerned with adaptation, but with comparing

exposed with non-exposed cases. Mann (2003) has conveniently summarised the design of

case–control studies (Fig. 4) and their properties:

Fig. 4 Schematic of the two-step case–control methodology employed in this study. When applied to pastdisaster studies, exposure or perturbation corresponds to the impact of a given hazard event, outcomecorresponds to the subsequent effects of a given event

Nat Hazards (2014) 71:335–362 345

123

1. They retrospectively compare two or more groups.

2. They aim to identify predictors of a given outcome.

3. They permit the assessment of the influence of predictors on the outcome of interest.

4. They are useful for hypothesis generation in relation to the underlying causal

mechanism or mechanisms.

He further notes that they are particularly useful for investigating infrequent events and

as such arguably constitute the method of choice for comparatively examining the impact

(= outcome) of natural hazards (= perturbation or exposure) on past communities.

Applying the case–control study design to historical or archaeological problems does,

however, come with a series of additional considerations. As Diamond and Robinson

(2010a) outline, the societies, cultures, or communities chosen for comparison can vary in

both their initial conditions, in the kind of exposure or perturbation, as well as in the

resulting outcomes. In addition, it is often difficult to assemble meaningful datasets suf-

ficiently large for the kind of extensive statistical manipulations routinely conducted by

epidemiologists. Finally, historic/archaeological case–control studies must additionally

account for the traditional weaknesses inherent in the method such as confounding vari-

ables and sampling bias. Yet as Diamond and Robinson also point out, these difficulties do

not in a major way detract from the utility of the case–control methodology and the rigour

it calls for in conducting comparative analyses. Even if historical scientists can never fully

satisfy the entire list of methodological requirements, formal comparative methods

nonetheless remain a fruitful avenue for retrospectively inferring general causality in

processes that unfold over time.

2.4 Comparative analysis: step 1

The evident cultural homogeneity of the hunter-gatherer communities distributed across

northern Europe prior to the Laacher See eruption suggests that differences in material

culture can be treated as a constant variable for the sake of this analysis. The different

regions under consideration in this study do, however, vary with respect to a series of other

parameters. These can be divided into geographic, ecological/economic, and demographic/

social variables, which, following the lead of Rolett and Diamond (2004), are rank-coded

(see Table 4). The outcome variable Impact is here scored a 0 = no impact, 1 = minor

cultural change, 2 = major cultural change, and 3 = abandonment. Likewise, the pertur-

bation variable (Tephra received) is coded according to how severe each region has been

affected by the fallout tephra, where 0 = no fallout tephra, 1 = thin distal cover,

2 = thicker medial fallout, and 3 = massive proximal deposits.

An initial bivariate Spearman’s correlation analysis does not reveal a statistically sig-

nificant link between the perturbation in the form of tephra fallout and the cultural impacts

reflected in the archaeological record. It does, however, reveal several significant corre-

lations between other variables (Table 5). The correlation between Distance from vent and

Time since colonisation is likely the result of confounding. In contrast, statistical signifi-

cance that can also be interpreted as substantive significance here links Latitude to

Resource diversity, Time since colonisation, and community Network position. Time since

colonisation is in turn significantly correlated with both Network position and Population

density. The correlation between the latter two variables fails to achieve significance, albeit

by only a very small margin. These correlations underline the overall ecological sensitivity

of these traditional societies and that the human re-colonisation of Europe can broadly be

346 Nat Hazards (2014) 71:335–362

123

Ta

ble

5R

esult

sof

atw

o-t

aile

dS

pea

rman

’sra

nk

corr

elat

ion

exer

cise

on

the

mat

rix

pre

sente

din

Tab

le4

Var

iab

lety

pes

Ou

tco

me

Per

turb

atio

nV

aria

ble

typ

es

Geo

gra

phic

Eco

logic

al/e

conom

icD

emogra

phic

/soci

al

Imp

act

Tep

hra

rece

ived

Lat

itud

eD

ista

nce

from

ven

tR

eso

urc

ed

iver

sity

Tim

esi

nce

colo

nis

atio

nN

etw

ork

po

siti

on

Po

pu

lati

on

den

sity

Co

rrel

ati

on

coef

fici

ent/

pva

lue

Imp

act

1

.

Tep

hra

rece

ived

0.5

57

1

0.1

94

.

Lat

itud

e0

.198

-0

.26

81

0.6

70

0.5

61

.

Dis

tan

cefr

om

ven

t0

.042

-0

.72

50

.46

21

0.9

29

0.0

65

0.2

96

.

Res

ou

rce

div

ersi

ty0

.076

0.1

59

-0

.87

4-

0.2

42

1

0.8

72

0.7

33

0.0

10

0.6

02

.

Tim

esi

nce

colo

nis

atio

n0

.198

-0

.58

60

.85

50

.80

4-

0.5

83

1

0.6

70

0.1

66

0.0

14

0.0

29

0.1

70

.

Net

wo

rkp

osi

tio

n0

.227

-0

.39

80

.87

40

.64

4-

0.7

50

0.8

74

1

0.6

25

0.3

76

0.0

10

0.1

18

0.0

52

0.0

10

.

Po

pu

lati

on

Den

sity

-0

.605

0.1

59

-0

.65

5-

0.7

25

0.4

17

-0

.80

1-

0.7

50

1

0.1

50

0.7

33

0.1

10

0.0

65

0.3

52

0.0

30

0.0

52

.

Co

rrel

atio

ns

sig

nifi

can

tat

pB

0.0

5ar

eg

iven

inb

old

Nat Hazards (2014) 71:335–362 347

123

understood as part of the more general ecological succession of plants and animals that

began in south-western Europe and was initiated by late ice age warming (Riede in press).

It is interesting to note that the Impact outcome of depopulation/observed cultural

changes is not significantly correlated with any of the chosen variables, although it is also

worth pointing out that trends towards significance can be observed between the outcome

and perturbation variables (p = 0.194) as well as between the outcome variable and

population density (p = 0.15). Principle components analysis is employed in order to

visualise and further scrutinise the relationships between the chosen variables. The first

three components explain [ 95 % of the variation in the present dataset, and Fig. 5 plots

the chosen variables in this three-component space. The outcome variable Impact and the

perturbation variable Tephra received plot tightly on component 2, whereas the relation-

ship between Impact and Distance from vent does not seem to be strong. The clustering of

variables on components 1 and 3, respectively, is ecological determined, i.e. the demog-

raphy and social landscape of expanding human communities in Europe appears to be

structured by the baseline of Resource diversity and Latitude, which pull the social/

demographic variables towards the extremes of their respective components.

In combination, these variables contribute to regionally differentiated vulnerabilities

that are the result of coupled ecological and social parameters. The ‘‘progression of vul-

nerability’’ (Wisner et al. 2004: 87) in relation to this calamity can be plotted much in the

same way as it can for more recent events, including the intersecting determinants of root

causes, dynamic pressures, and unsafe conditions that, together with the hazard in question,

resulted in marked societal changes (Fig. 6). The effects of this eruption include the

emergence of the Bromme culture in southern Scandinavia and the possible abandonment

of areas to the northeast of the eruptive centre. These effects can be traced over decadal and

centennial timescales. More speculatively, it can be added that the form of projectile points

that came to dominate in southern Scandinavia remains popular in the region for many

centuries and indeed millennia, albeit in a miniaturised form. It is not possible to say

whether these elements of material culture were also linked to specific traditional

knowledge regarding the eruption, but the creation and maintenance of such encoded

information is—in analogy with similar cases from the ethno-historic record elsewhere

(Beaudoin and Oetelaar 2006; Blong 1982; Vitaliano 2007)—likely.

Fig. 5 A visualisation of crudesocio-ecological vulnerability indifferent regions of Europe at theend of the last ice age usingprinciple components analysis

348 Nat Hazards (2014) 71:335–362

123

2.5 Comparative analysis: step 2

Can the detailed synchronic observations of the Laacher See eruption’s contemporaneous

impacts be generalised beyond its time and across societies of different socio-economic

integration? Much later in prehistory, Europe was again affected by the far-field impact of

a volcanic eruption. In one of the years around 1610 BCE, the Thera volcano, today better

know as the popular Aegean island of Santorini, erupted. Although the dating of this event

has been contentious, recent efforts have been able to quite precisely pinpoint it to the

period between 1627 and 1600 BCE (Warburton et al. 2009; Friedrich et al. 2006). The

island, which at this time was a thriving and important political and economic centre in the

Bronze Age Mediterranean world, was devastated (Renfrew 1979). The effects of this

eruption can be traced in the economic as well as the cultural spheres (Driessen and

MacDonald 2000, 1997; Bicknell 2000). In fact, its destruction may have precipitated the

long-lived Atlantis myth (Friedrich 2009). Beyond the Thera eruption’s immediate impact

on island communities, its effects reverberated through the complex networks of economic

and diplomatic relations around the Mediterranean at that time, leading eventually to major

political reconfigurations (Knappett et al. 2011). Furthermore, the indirect effects of this

eruption can also arguably be traced in Europe north of the Alps, where the collapse of

southern trade connections left communities more vulnerable to the general climatic effects

of the eruption (Baillie 1991).

In or just before 536 CEVolcan Ilopango in modern-day El Salavador erupted. This very

large eruption is now dated with great precision in both terrestrial and ice-core records

Fig. 6 A pressure and release (PAR) schematic for the impact of the Laacher See eruption oncontemporaneous communities. IACP = Intra-Allerød Cold Phase, a 200–300-year episode of colder andmore variable climate within the otherwise relatively warm and stable so-called Allerød period. The LaacherSee volcano erupted in the latter part of this period just following the IACP. The magnifying glass indicatesfocus on the community scale, where vulnerability and resilience are shaped by access to resources and localpolitical conditions. Redrawn from Wisner et al. (2004)

Nat Hazards (2014) 71:335–362 349

123

(Mehringer et al. 2005; Larsen et al. 2008). Its impact in the near-field on Maya farming

communities in Central America has long been discussed (Sheets 2008, 2007, 2001; Dull

et al. 2001; Sheets 1999, 1981, 1979), but it is only recently that an appreciation of its

potential far-field impact is growing, both globally as well as specifically in Europe. A

widespread horizon of societal change, the so-called ‘‘AD 536 event’’, has been recognised

around the world (Gunn 2000; Baillie 1991) and given recent improvements in dating of

the Ilopango event (e.g. Dull et al. 2001; Mehringer et al. 2005; Larsen et al. 2008) it is

looking increasingly likely that the eruption can be linked to it. In northern Europe, the

‘‘AD 536 event’’ takes the form of a profound series of cascading demographic, political,

and religious changes that are reflected in both settlement patterns, economy, religious

behaviours, and even in myth-making amongst complex early Medieval state societies,

elements of which (e.g. the so-called Fimbul Winter of Nordic mythology) survive to the

present day (Graslund and Price 2012; Graslund 2008; Lowenborg 2012; Arrhenius 2013;

Axboe 1999, 2001). Unaware of the original source of their troubles, these communities

were affected only by the secondary meteorological and climatic aftermath of the Ilopango

event, but such effects are known from more recent eruptions (e.g. Tambora), which have

impacted people in both definite economic terms but also in relation to their world view

(Stommel and Stommel 1983; Kramer 2009; Oppenheimer 2003).

More recently still, the 2010 eruption of Eyjafjallajokull (Iceland) also impacted Europe

in its far-field. This event—geologically speaking ‘‘small-scale’’ and ‘‘rather ordinary’’

(Davies et al. 2010: 606 and 608)—evolved from an initially not at all unwelcome tourist

attraction (Benediktsson et al. 2011) into what eventually was at least perceived by many

as a major disaster (Lund and Benediktsson 2011) that had far-reaching and economically

severe effects (Pedersen 2010). The differences between near-field impacts on plants,

animals, people, and landscape (e.g. Bird and Gısladottir 2012; Bird et al. 2011) and the

far-field impacts on patterns of mobility, economics, and technology (Birtchnell and

Buscher 2011) varied dramatically and illustratively. Only comparatively few people were

affected proximally, and local communities as well as state services were well prepared

offering financial and engineering assistance. Yet amongst those most directly affected by

the eruption, remarkable new material culture patterns and behaviours can be recognised

(see, for instance, http://www.icelanderupts.is/). Distally, the effects of the eruption

reverberated through and were aggravated by the networks of ‘‘fragile mobilities’’ (Lund

and Benediktsson 2011: 8) it impacted and can be measured in considerable economic

losses as well as individual nuisance.

It is too early to judge which if any long-term impacts this eruption caused. Yet the

commonalities flagged up in all the case studies invoked here—first and foremost the

network sensitivity of the affected communities, the almost universal mobility-related

responses, and the transformation and occasionally amplification of impacts in the far-

field—do have implications for disasters planning in the present day. Returning to the

Laacher See case study, it should be noted that the Eifel volcanic system is by no means

extinct. The mantle plume underneath the Laacher See remains active even if currently

dormant (Ritter et al. 2001; Zhu et al. 2012) and with volcanic activity in the area possibly

linked to periods of warm climates (Nowell et al. 2006), recent anthropogenic temperature

rises may initiate renewed eruptions. Park and Schmincke (1997: 523) warn: ‘‘Recurrence

of a major eruption of LSV (Laacher See volcano) would no doubt generate phreato-

magmatic explosions more powerful than those 12,900 years ago and would pose a major

hazard and risk to the densely populated and highly industrialised lowland of Neuwied

Basin’’. As can be seen from the effects of the past eruptions discussed here, the effects of

such an eruption would extend well beyond this immediate area. In addition, these impacts

350 Nat Hazards (2014) 71:335–362

123

may be further aggravated during times of economic vicissitude or political tensions to

which even Europe, an otherwise economically and politically stable region, is not

immune.

3 Discussion

A formal case–control study of the impact of the high-magnitude mid-continent prehistoric

Laacher See volcanic eruption has been attempted. Sample size is painfully small and the

archaeological perspective on past communities woefully coarse. The lack of a clear

association between perturbation and outcome reveals a complex causal mosaic of how this

eruption affected contemporaneous communities. Can we nonetheless draw some useful

conclusions from such a comparative analysis? Vulnerability is widely understood as a

complex ‘‘multidimensional’’ property (Yoon 2012: 824) that combines ecological and

social parameters (Cutter et al. 2003; Cutter 1996; Birkmann et al. 2013). Multivariate data

analysis methods are well suited for exploring such multidimensionality, and principle

components analysis suggests that strong geographic and ecological gradients structured

the access to social and physical resources amongst late ice age hunter-gatherers. This

‘‘spatiality of risk’’ (November 2008: 1523) played a more important role at that time than

the crude amount of fallout tephra received. Importantly, distance from the eruptive centre

also does not appear to be decisive in relation to post-event impacts. Population density is

argued to drive vulnerability, with lower population levels in this case leading to height-

ened sensitivity to extreme event disturbances. At lower latitudes and amongst societies

following different economic strategies, the relationship between population density and

vulnerability may have been quite different, however (see Sheets 2008, 2012). In the

present day, population growth and urban clumping are without doubt again leading to

heightened levels of vulnerability (e.g. Small and Naumann 2001), suggesting that there

may exist a demographic middle ground of optimal resilience or that relatively devolved

and decentralised demographic structures strengthen resilience.

In all case studies examined social networks likewise have vital ramifications for

understanding causalities, especially as effects in the far-field are not readily predicted by

the geophysical parameters of the event in question. The more complex the affected

network structures, the more unpredictable and potentially cascading the effects of a given

eruption can be, a finding in line with previous suggestions (e.g. Sheets 2008; Knappett

et al. 2011). Similarly, volcanic eruptions appear to impact mobility, either by restricting it

or by leading to migration. Interestingly, population displacement is also by far the most

common effect of volcanic eruptions in recent times (Witham 2005), suggesting that

movement away from actual or perceived hazards is a major coping strategy adopted by

individuals and groups at many if not all levels of socio-economic integration.

One of the strengths of the case–control methodology is to suggest further hypotheses

about the mechanisms linking perturbation and outcome. In the case of the Laacher See

event, some hypotheses have already been evaluated (Riede and Bazely 2009; Riede and

Wheeler 2009), others (e.g. the effects of the ash’s chemical loading) remain to be

explored in detail. Each of these additional explorations has further strengthened the con-

clusion that the eruption’s impact was as pronounced or even more so in the far-field as in

the near-field, but that the mechanisms of impact changed along the proximal-to-distal axis.

The occurrence of a high-magnitude volcanic eruption from, for instance, the Laacher

See edifice in the heart of Europe is currently only entertained in fiction (Schreiber 2006)

and the tabloid press (see, for example, http://www.bild.de/news/2007/news/forscher-

Nat Hazards (2014) 71:335–362 351

123

ausbruch-deutschland-1399914.bild.html and the discussion at http://www.wired.com/

wiredscience/2012/01/fearmongering-gets-started-in-2012-laacher-see-is-not-ready-to-

blow/), but Clarke (2006) has long argued that the potential impact of such extreme

events—no matter how improbable—should be countenanced in order to evaluate the

‘‘surge capacity’’ of emergency systems (Clarke 2008a: 638) and thereby to increase both

event-specific but also general resilience in the present day (see also Michel-Kerjan 2012).

He argues that one tool for pondering such ‘‘worst cases’’ is counterfactual reasoning

(Clarke 1999, 2008a, b). Complementary to this, the detailed information available from

past eruptions such as the 11,000 BCE Laacher See event can provide important clues for

‘‘retrofactually’’ considering the impact that renewed volcanic activity in the Eifel would

have locally, regionally, as well as superregionally.

If we triangulate between the four case studies from the deep past to the present day,

much more robust scenarios of not just the possible but the likely impact of a future

eruption of the Laacher See volcano can be derived. The full suite of mechanisms now

known to link LST deposition to cultural consequences via their attendant impacts on

ecosystems would also be relevant in future events of this kind. Furthermore, all case

studies discussed here highlight the critical nature of communication channels, of social

networks, and of mobility. Self (2006) calculates that the probability of a M = 6 eruption

such as the LSE to occur in the twenty-first century to be 100 %, and he tersely describes

some of the likely consequences of an eruption like this or larger. Although by no means

the most probable candidate, such a low-frequency/high-magnitude mid-continent eruption

in Europe lasting several weeks or months would likely lead to a prolonged closure of

European or even Eurasian airspace, an at least temporary collapse of air- and water-based

supply chains providing many daily consumables, and key power supply nodes would be at

risk. The economic implications of these immediate effects and their longer-term cleanup/

repair efforts are staggering and would likely put European economic as well as political

systems under considerable strain. Migration, political, but also religious changes were

demonstrably the results of such eruptions in the past and must be taken serious as potential

effects of future eruptions (Chester 2005). Fortunately, given the very low probability of

renewed activity at the Laacher See in the near future, the scenario sketched out here is no

more than a thought experiment. By their very nature, however, the far-field effects such as

those described here could also occur as the result of eruptions at the volcanically highly

active European periphery (i.e. Iceland or the Mediterranean) or even further away.

4 Conclusion

The frequency of extreme events is predicted to rise in the future (Field et al. 2012; Hoyois

and Guha-Sapir 2012). Although volcanic eruptions do not rank highly amongst the most

lethal of geological events, they have often lead to widespread homelessness and large-

scale migration in recent history as well as in the deep past. Uncontrolled migration,

economic, political, and religious destabilisation and radicalisation are seen as major

threats to society (e.g. Howell 2013). Migration, in particular, has been identified as a

significant political and logistic challenge (Black et al. 2011). The case studies presented

here as well as other less formal comparative investigations of past volcanism (Sheets

2001, 2012; Ort et al. 2008) support the view that also in the long-term volcanic eruption

can have significant, far-reaching, and lasting societal effects often via precisely such

indirect mechanisms. Future volcanic crises will likely be aggravated by globally

increasing populations (Small and Naumann 2001), the preferential clustering of, in

352 Nat Hazards (2014) 71:335–362

123

particular, the urban poor in zones of volcanic risk (Chester et al. 2001), as well as the high

network interdependence of contemporary economies (Helbing 2013; Bailey 2011). The

prospect of these coupled risk and vulnerability parameters makes it an urgent matter to

better understand volcanic events and their direct and indirect societal impacts.

Hulme (2008: 5) has noted that ‘‘we are living in a climate of fear about our future

climate. The language of the public discourse around global warming routinely uses a

repertoire which includes words such as ‘‘catastrophe’’, ‘‘terror’’, ‘‘danger’’, ‘‘extinction’’,

and ‘‘collapse’’. To help make sense of this phenomenon, the story of the complex rela-

tionships between climates and cultures in different times and in different places is in

urgent need of telling. If we can understand from the past something of this complex

interweaving of our ideas of climate with their physical and cultural settings, we may be

better placed to prepare for different configurations of this relationship in the future’’.

Increasingly, archaeological datasets are being used to this end (e.g. Gerrard and Petley

2013; McCormick et al. 2007; Mitchell 2008; Van de Noort 2011), and this paper has

preliminarily attempted to further position social, ecological, and place vulnerability of

traditional prehistoric societies to volcanic eruptions in a long-term perspective. A two-step

methodology has been presented where first a deep time case study is analysed in a formal

comparative manner and then set into a less constrained comparative perspective that

bridges pre-industrial and industrial societal forms. Important data weaknesses remain, and

several methodological hurdles need to be overcome if the comparative study of past

disasters is to make a substantive contribution to Disaster Risk Reduction research. For

instance, issues of variable impact amongst different social classes or genders should be

addressed whenever possible. In addition, both geographic and temporal scales are

important when comparing vulnerability and impact across past disasters, just as they are

important when assessing vulnerability in the present (Fekete et al. 2010). Nonetheless, the

commonalities and differences in vulnerability revealed by diachronic and cross-cultural

analysis provide pointers for building post-industrial comprehensive resilience: Such

resilience has a strong geographic and demographic dimension, must engage social

structure, and political as well as religious concerns.

A number of excellent databases covering Quaternary volcanism exist (e.g. Bryson et al.

2006; Crosweller et al. 2012; http://www.volcano.si.edu/). These data repositories gener-

ally contain little or no information on affected societies, however, and are thus of limited

use when the aim is to study vulnerability and societal impacts. In contrast to such large

datasets, many archaeological or historical studies of past disasters suffer from being

overly particularistic. Numerous excellent collections of such case studies exist (specifi-

cally for volcanic eruption see, for example, Grattan and Torrence 2007a; Torrence and

Grattan 2002; McGuire et al. 2000; Raynal et al. 2002; Oppenheimer 2011; de Boer and

Sanders 2002; McCoy and Heiken 2000), but they arguably do better at showcasing the

diversity of volcanic events and affected societies than at comparing these events. Like-

wise, disaster scientists working in the present day are hampered by the fact that they study

seemingly unique events, which usually have long recurrence intervals when measured on

a human timescale. This restricted event database makes formal, quantitative, or indeed

qualitative comparative analyses difficult. In order to work towards uniting these diverging

approaches, this paper has focused deliberately on cases that share important analytical

variables, so that differences between cases are minimised. The natural experimental

methodology employed here in turn facilitates insights into causality. Whilst such formal

analyses do not replace detailed descriptive case research—indeed they rely on it—this

paper suggests that more powerful scenarios can eventually be derived through a mixed-

method approach that combines case-based and comparative angles.

Nat Hazards (2014) 71:335–362 353

123

As Alexander (1997: 289) has noted, ‘‘every natural disaster involves a unique pattern

of physical energy expenditure and human reaction. Nevertheless, there is sufficient

similarity between events to enable one to distinguish common phases of the emergency,

typical responses, and characteristic patterns of impact…Generality is clearly more

important than uniqueness in characterising disaster’’. The inclusion of past calamities

would allow a notable increase in the number of event samples, which in turn opens the

possibility for more robust comparative analyses in the search for important generalities.

Owing to the vicissitudes of data resolution, such analyses must focus on the kinds of

information readily accessible in the historical and archaeological records, i.e. basic

demographic and economic patterns, the built environment, social organisation, and reli-

gious beliefs as reflected in material culture. By the same token, a re-orientation towards

past events and societies leads to a concomitant recasting of vulnerability in light of these

societal features, which in turn could be targeted when insights from such studies should be

implemented to reduce contemporary vulnerability amongst at-risk communities. Given

the dramatic differences between past and present communities in Europe and elsewhere,

we cannot hope to gain insights into how to improve resilience from a management

perspective (Toft and Reynolds 1994). Instead, an archaeological and historical perspective

on calamities provides information at the community level and can reveal non-trivial

aspects of vulnerability that are difficult to grasp otherwise. The collective cultural heritage

of the archaeological and historical records may allow a retrieval or reconstruction of past

local knowledge (Hilhorst and Bankoff 2004) that may help local communities to both

prepare for and cope with future calamities (Cashman and Cronin 2008; Chester and

Duncan 2007). Archaeological and historical data can thus provide immediacy to hazard

forecast scenarios (e.g. Mastrolorenzo et al. 2006), and these disciplines’ already estab-

lished channels of communication (i.e. museum exhibitions) could be used to disseminate

scientific results and mitigation plans at the community level. Including data from disci-

plines with both short- and long-term perspectives is particularly relevant when repose

times between hazard events are long. In this way, the rich data provided by these elements

of cultural heritage—our ‘‘usable past’’ (Stump 2013: 268)—could play a more proactive

role in present and future risk reduction strategies and in the strengthening of social

resilience that emerges out of a coupling between traditional and scientific knowledge and

methods (Dix and Rohrs 2007; Lorenz 2013; Donovan et al. 2011). Finally, it should be

noted that in the past (Grattan and Torrence 2007b) as well as in the present (Birkmann

et al. 2010; Olshansky et al. 2012), extreme events and their societal impacts in principle

also offer opportunities for creative and accelerated positive culture change rooted in

aspects of resilience.

In summary, this paper has argued that the Laacher See eruption acted as both a

‘‘trigger’’ and ‘‘catalyst’’ (Garcıa-Acosta 2002: 57) of social change in eleventh millennium

BCE Europe. In turn, the eruptions of the Laacher See volcano, of Thera, of Volcan

Ilopango, and of Eyjafjallajokull have together functioned as ‘‘revealers…of preexisting

critical conditions’’ (Garcıa-Acosta 2002: 57) in the form of quite specific systemic

weaknesses and vulnerabilities amongst European communities at each time slice. These

weaknesses have implications for thinking about future extreme events not just in this

region but also elsewhere. The aim of this paper has not been to reinvent ‘‘the wheel of

‘disasterology’’’ (Alexander 1997: 298), but to draw the outline of a research programme

combining Risk Reduction Research, archaeology, and volcanology. Importantly, by

studying past societal impacts, we can derive historically informed evidence-based policy

recommendations that can provide otherwise purely symbolic planning with operational

and functional dimensions (sensu Clarke 1999) and that can be part of culturally sensitive

354 Nat Hazards (2014) 71:335–362

123

social resilience strategies (sensu Lorenz 2013). The methodology of contemporary

disaster science can inform how we investigate the impact of extreme events on past

communities. In return, the methodology for a science of past disasters and the case studies

discussed in this paper do offer a genuine opportunity for extending disaster science into

the deep past and thereby strengthening the discipline as a whole.

Acknowledgments The Laboratory of Past Disaster Science (LaPaDiS) is a novel collaborative effort byhistorical, social, and natural scientists anchored at Aarhus University. LaPaDiS is generously funded by theDanish Agency for Science, Technology and Innovation Grant No. 11-106336. The critical reading andcomments of three anonymous reviewers are greatly appreciated; each has contributed to an improvedmanuscript.

References

Adger WN (2006) Vulnerability. Glob Environ Chang 16(3):268–281Alexander DE (1995) A survey of the field of natural hazards and disaster studies. In: Carrara A, Guzetti F

(eds) Geographical information systems in assessing natural hazards. Kluwer, Amsterdam, pp 1–19Alexander DE (1997) The study of natural disasters, 1977–1997: some reflection on a changing field of

knowledge. Disasters 21(4):284–304Alvarez-Fernandez E (2009) Magdalenian personal ornaments on the move: a review of the current evidence

in Central Europe. Zephyrus 63:45–59Arrhenius B (2013) Helgo in the shadow of the dust veil 536–537. J Archaeol Anc Hist 5:1–14Axboe M (1999) The year 536 and the Scandinavian gold hoards. Mediev Archaeol 43:186–188Axboe M (2001) Amulet pendants and a darkened sun. In: Magnus B (ed) Roman gold and the development

of the early Germanic kingdoms, history and antiquities. Royal Academy of Letters, Stockholm,pp 119–136

Baales M (2002) Der spatpalaolithische Fundplatz Kettig: Untersuchungen zur Siedlungsarchaologie derFedermesser-Gruppen am Mittelrhein. Verlag Rudolf Habelt GmbH, Bonn

Baales M, Joris O, Street M, Bittmann F, Weninger B, Wiethold J (2002) Impact of the late glacial eruptionof the Laacher See Volcano, Central Rhineland, Germany. Quat Res 58(3):273–288

Bailey AJ (2011) Population geographies and climate change. Prog Hum Geog 35(5):686–695Baillie MGL (1991) Marking in marker dates: towards an Archaeology with historical precision. World

Archaeol 23(2):233–243Bandi H-G (1968) Das Jungpalaolithikum. Ur- und Fruhgeschichtliche Archaologie der Schweiz 1:107–122Bankoff G (2004) Time is of the essence: disasters, vulnerability and history. Int J Mass Emerg Disasters

22(3):23–42Beaudoin AB, Oetelaar GA (2006) The day the dry snow fell: the record of a 7627-year-old disaster. In:

Payne M, Wetherell D, Kavanaugh C (eds) Alberta formed Alberta transformed, vol 1. University ofAlberta Press/University of Calgary Press, Edmonton

Benediktsson K, Lund KA, Huijbens E (2011) Inspired by eruptions? Eyjafjallajokull and Icelandic tourism.Mobilities 6(1):77–84

Bicknell P (2000) Late Minoan IB marine ware, the marine environment of the Aegean, and the Bronze Ageeruption of the Thera volcano. Geol Soc Spec Publ 171(1):95–103

Bird D, Gısladottir G (2012) Residents’ attitudes and behaviour before and after the 2010 Eyjafjallajokulleruptions—a case study from southern Iceland. Bull Volcanol 74(6):1263–1279

Bird D, Gısladottir G, Dominey-Howes D (2011) Different communities, different perspectives: issuesaffecting residents’ response to a volcanic eruption in southern Iceland. Bull Volcanol 73(9):1209–1227

Birkmann J, Buckle P, Jaeger J, Pelling M, Setiadi N, Garschagen M, Fernando N, Kropp J (2010) Extremeevents and disasters: a window of opportunity for change? Analysis of organizational, institutional andpolitical changes, formal and informal responses after mega-disasters. Nat Hazards 55(3):637–655

Birkmann J, Cardona OD, Carreno ML, Barbat AH, Pelling M, Schneiderbauer S, Kienberger S, Keiler M,Alexander D, Zeil P, Welle T (2013) Framing vulnerability, risk and societal responses: the MOVEframework. Nat Hazards 67(2):193–211

Birks HJB, Lotter AF (1994) The impact of the Laacher See Volcano (11000 year B.P.) on terrestrialvegetation and diatoms. J Paleolimnol 11:313–322

Birtchnell T, Buscher M (2011) Stranded: an eruption of disruption. Mobilities 6(1):1–9

Nat Hazards (2014) 71:335–362 355

123

Black R, Bennett SRG, Thomas SM, Beddington JR (2011) Climate change: migration as adaptation. Nature478(7370):447–449

Blong RJ (1982) The time of darkness. Local legends and volcanic reality in Papua New Guinea. Universityof Washington Press, Seattle

Bocquet-Appel J-P, Demars P-Y, Noiret L, Dobrowsky D (2005) Estimates of Upper Palaeolithic meta-population size in Europe from archaeological data. J Archaeol Sci 32:1656–1668

Bodu P (1998) Magdalenians-early Azilians in the centre of the Paris Basin: a filiation? The example of LeCloseau (Rueil-Malmaison, France). In: Milliken S (ed) The organization of lithic technology in lateglacial and early postglacial of Europe. Oxbow, Oxford, pp 131–147

Brauer A, Endres C, Negendank JFW (1999) Lateglacial calendar year chronology based on annuallylaminated sediments from Lake Meerfelder Maar, Germany. Quat Int 61:17–25

Brinch Petersen E (2009) The human settlement of southern Scandinavia 12500–8700 cal BC. In: Street M,Barton RNE, Terberger T (eds) Humans, environment and chronology of the late glacial of the NorthEuropean Plain. RGZM—Tagungen, Band 6. Verlag des Romisch-Germanischen Zentralmuseums,Mainz, pp 89–129

Bryson RU, Bryson RA, Ruter A (2006) A calibrated radiocarbon database of late quaternary volcaniceruptions. eEarth Discussions 1:123–134

Cashman KV, Cronin SJ (2008) Welcoming a monster to the world: myths, oral tradition, and modernsocietal response to volcanic disasters. J Volcanol Geoth Res 176(3):407–418

Chester DK (2005) Theology and disaster studies: the need for dialogue. J Volcanol Geoth Res146(4):319–328

Chester DK, Duncan AM (2007) Geomythology, theodicy, and the continuing relevance of religiousworldviews on responses to volcanic Eruptions. In: Grattan J, Torrence R (eds) Living under theshadow. Cultural impacts of volcanic eruptions. Left Coast Press, Walnut Creek, pp 203–224

Chester DK, Degg M, Duncan AM, Guest JE (2001) The increasing exposure of cities to the effects ofvolcanic eruptions: a global survey. Environ Hazards 2(3):89–103

Chester DK, Duncan AM, Sangster H (2012) Human responses to eruptions of Etna (Sicily) during the late-pre-industrial era and their implications for present-day disaster planning. J Volcanol Geoth Res225–226:65–80

Clarke L (1999) Mission improbable. Using fantasy documents to tame disaster. The University of ChicagoPress, Chicago

Clarke L (2006) Worst cases. Terror and catastrophe in the popular imagination. The University of ChicagoPress. Chicago

Clarke L (2007) Thinking possibilistically in a probabilistic world. Significance 4(4):190–192Clarke L (2008a) Possibilistic thinking: a new conceptual tool for thinking about extreme events. Soc Res

75(3):669–690, 1033Clarke L (2008b) Thinking about worst-case thinking. Sociol Inq 78(2):154–161Cooper J, Sheets PD (eds) (2012) Surviving sudden environmental change. University of Colorado Press,

BoulderCoudret P, Fagnart J-P (1997) Les industries a Federmesser dans le bassin de la Somme: chronologie et

identite des groupes culturels. Bulletin de la Societe Prehistorique Francaise 94(3):349–360Cronin S, Petterson M, Taylor P, Biliki R (2004a) Maximising multi-stakeholder participation in govern-

ment and community volcanic hazard management programs; a case study from Savo, SolomonIslands. Nat Hazards 33(1):105–136

Cronin SJ, Gaylord DR, Charley D, Alloway BV, Wallez S, Esau JW (2004b) Participatory methods ofincorporating scientific with traditional knowledge for volcanic hazard management on Ambae Island,Vanuatu. Bull Volcanol 66(7):652–668

Crosweller HS, Arora B, Brown SK, Cottrell E, Deligne NI, Guerrero NO, Hobbs L, Kiyosugi K, LoughlinSC, Lowndes J, Nayembil M, Siebert L, Sparks RSJ, Takarada S, Venzke E (2012) Global database onlarge magnitude explosive volcanic eruptions (LaMEVE). J Appl Volcanol 1(1):4

Cutter SL (1996) Vulnerability to environmental hazards. Prog Hum Geog 20(4):529–539Cutter SL, Boruff BJ, Shirley WL (2003) Social vulnerability to environmental hazards. Soc Sci Quart

84(2):242–261Davies SM, Larsen G, Wastegard S, Turney CSM, Hall VA, Coyle L, Thordarson T (2010) Widespread

dispersal of Icelandic tephra: how does the Eyjafjoll eruption of 2010 compare to past Icelandicevents? J Quat Sci 25(5):605–611

De Belizal E, Lavigne F, Gaillard JC, Grancher D, Pratomo I, Komorowski J-C (2012) The 2007 eruption ofKelut volcano (East Java, Indonesia): phenomenology, crisis management and social response. Geo-morphology 136(1):165–175

356 Nat Hazards (2014) 71:335–362

123

De Bie M, Caspar J-P (2000) Rekem. A Federmesser Camp on the Meuse River Bank, vol 1. LeuvenUniversity Press, Leuven

De Bie M, Vermeersch PM (1998) Pleistocene-Holocene transition in the Benelux. Quat Int 49(50):29–43de Boer JZ, Sanders DT (2002) Volcanoes in human history. The far-reaching effects of major eruptions.

Princeton University Press, Princetonde Klerk P, Janke W, Kuhn P, Theuerkauf M (2008) Environmental impact of the Laacher See eruption at a

large distance from the volcano: integrated palaeoecological studies from Vorpommern (NE Ger-many). Palaeogeogr Palaeocl 270(1–2):196–214

Debout G, Olive M, Bignon O, Bodu P, Chehmana L, Valentin B (2012) The Magdalenian in the ParisBasin: new results. Quat Int 272–273:176–190

Dev S, Riede F (2012) Quantitative functional analysis of late glacial projectile points from northernEurope. Lithics 33:40–55

Diamond JM, Robinson JA (2010a) Afterword: using comparative methods in studies of human history. In:Diamond JM, Robinson JA (eds) Natural experiments of history. Belknap Press, Cambridge,pp 267–275

Diamond JM, Robinson JA (eds) (2010b) Natural experiments of history. Belknap Press, CambridgeDix A, Rohrs M (2007) Vergangenheit versus Gegenwart? Anmerkungen zu Potentialen, Risiken und

Nebenwirkungen einer Kombination historischer und aktueller Ansatze der Naturgefahrenforschung.Hist Soc Res 32(3):215–234

Donovan A, Oppenheimer C, Bravo M (2011) Rationalising a volcanic crisis through literature: Mont-serratian verse and the descriptive reconstruction of an island. J Volcanol Geoth Res 203(3–4):87–101

Driessen J, MacDonald CF (1997) The troubled Island: Minoan Crete before and after the Santorinieruption. Aegaeum No.17. Universite de Liege, Histoire de l’art et archeologie de la Grece antiqueLiege

Driessen J, MacDonald CF (2000) The eruption of the Santorini volcano and its effects on Minoan Crete.Geol Soc Spec Publ 171(1):81–93

Dull RA, Southon JR, Payson S (2001) Volcanism, ecology and culture: a reassessment of the VolcanIlopango Tbj eruption in the Southern Maya Realm. Lat Am Antiq 12(1):25–44

Dunning T (2008) Improving causal inference: strengths and limitations of natural experiments. Polit ResQuart 61(2):282–293

Dunning T (2012) Natural experiments in the social sciences. Cambridge University Press, CambridgeEriksen BV (1996) Resource Exploitation, Subsistence Strategies, and Adaptiveness in Late Pleistocene-

Early Holocene Northwest Europe. In: Straus LG, Eriksen BV, Erlandson JM, Yesner DR (eds)Humans at the end of the ice age. The archaeology of the pleistocene-holocene transition. PlenumPress, New York, pp 101–128

Erkens G, Hoffmann T, Gerlach R, Klostermann J (2011) Complex fluvial response to Lateglacial andHolocene allogenic forcing in the Lower Rhine Valley (Germany). Quat Sci Rev 30(5–6):611–627

Fekete A, Damm M, Birkmann J (2010) Scales as a challenge for vulnerability assessment. Nat Hazards55(3):729–747

Felgentreff C, Glade T (eds) (2008) Naturrisiken und Sozialkatastrophen. Spektrum Akademischer Verlag,Berlin

Field CB, Barros V, Stocker TF, Qin D, Dokken DJ, Ebi KL, Mastrandrea MD, Mach KJ, Plattner G-K,Allen SK, Tignor M, Midgley PM (eds) (2012) Managing the risks of extreme events and disasters toadvance climate change adaptation. Working groups I and II of the intergovernmental panel on climatechange, special report. Cambridge University Press, Cambridge

Fischer A, Mortensen MF, Henriksen PS, Mathiassen DR, Olsen J (2013) Dating the Trollesgave site and theBromme culture—chronological fix-points for the Lateglacial settlement of Southern Scandinavia.J Archaeol Sci 40(12):4663–4674

Fisher RV, Schmincke H-U (1984) Pyroclastic rocks. Springer, BerlinFitzhugh B (2012) Hazards, impacts, and resilience among hunter-gatherers of the Kuril Islands. In: Cooper

J, Sheets PD (eds) Surviving sudden environmental change. University of Colorado Press, Boulder,pp 19–42

Friedrich WL (2009) Santorini: volcano, natural history, mythology. Aarhus University Press, ArhusFriedrich WL, Kromer B, Friedrich M, Heinemeier J, Pfeiffer T, Talamo S (2006) Santorini eruption

radiocarbon dated to 1627–1600 BC. Science 312(5773):548Galavotti MC (2003) Observation and experiment in the natural and social sciences. Kluwer, AmsterdamGamble C, Davies W, Pettitt P, Richards M (2004) Climate change and evolving human diversity in Europe

during the last glacial. Philos Trans R Soc B 359:243–254

Nat Hazards (2014) 71:335–362 357

123

Garcıa-Acosta V (2002) Historical disaster research. In: Hoffman SM, Oliver-Smith A (eds) Catastrophe &culture: the anthropology of disaster. School of American research advanced seminar series. School ofAmerican Research Press, Santa Fe, pp 49–66

Gelhausen F, Kegler J, Wenzel S (2004) Hutten Oder Himmel? Latente Behausungsstrukturen imSpatpalaolithikum Mitteleuropas. Jahrbuch des Romisch-Germanischen Zentralmuseum Mainz51(1):1–22

Gerrard C, Petley D (2013) A risk society? Environmental hazards, risk and resilience in the later middleages in Europe. Nat Hazards 69(1):1051–1079

Graf H-F, Timmreck C (2001) A general climate model simulation of the aerosol radiative effects of theLaacher See eruption (10,900 BC). J Geophys Res 106(14):14747–14756

Graslund B (2008) Fimbulvintern, Ragnarok och klimatkrisen ar 536–537 e. Kr. Saga och Sed 2007:93–123Graslund B, Price N (2012) Twilight of the gods? The ‘dust veil event’ of AD 536 in critical perspective.

Antiquity 86(332):428–443Grattan J, Torrence R (eds) (2007a) Living under the shadow. Cultural impacts of volcanic eruptions. One

world archaeology no. 53. Left Coast Press, Walnut CreekGrattan J, Torrence R (2007b) Beyond gloom and doom: the long-term consequences of volcanic disasters.

In: Grattan J, Torrence R (eds) Living under the shadow. Cultural impacts of volcanic eruptions. LeftCoast Press, Walnut Creek, pp 1–18

Grayson DK, Sheets PD (1979) Volcanic disasters and the archaeological record. In: Sheets PD, GraysonDK (eds) Volcanic Activity and Human Ecology. Academic Press, London, pp 623–632

Grunberg JM (2006) New AMS dates for Palaeolithic and Mesolithic camp sites and single finds in Saxony-Anhalt and Thuringia (Germany). Proc Prehist Soc 72:95–112

Gunn JD (ed) (2000) The years without summer. Tracing AD 536 and its aftermath. British archaeologicalreports (international series) 872. Archaepress, Oxford

Harms E, Schmincke H-U (2000) Volatile composition of the phonolitic Laacher See magma (12,900 yearBP): implications for syn-eruptive degassing of S, F, Cl and H2O. Contrib Min Petr 138(1):84–98

Harvey PH, Pagel MD (1991) The comparative method in evolutionary biology. Oxford University Press,Oxford

Helbing D (2013) Globally networked risks and how to respond. Nature 497(7447):51–59Helbling J (2006) Coping with ‘natural’ disasters in pre-industrial societies: some comments. Mediev Hist J

10(1–2):429–446Hewitt K (ed) (1983) Interpretations of calamity. Allen & Unwin Inc., BostonHilhorst T, Bankoff G (2004) Introduction: mapping vulnerability. In: Hilhorst T, Frerks G, Bankoff G (eds)

Mapping vulnerability: disasters, development, and people. Earthscan Publications, London, pp 1–9Howell L (ed) (2013) Global risks 2013. World economic forum risks reports, 8th edn. World Economic

Forum, GenevaHoyois P, Guha-Sapir D (2012) Measuring the human and economic impact of disasters. CRED/Govern-

ment Office for Science, LouvainHulme M (2008) The conquering of climate: discourses of fear and their dissolution. Geogr J 174(1):5–16Janku A, Schenk GJ, Mauelshagen F (eds) (2012) Historical disasters in context: science, religion, and

politics. Routledge, New YorkJanssens MM, Kasse C, Bohncke SJP, Greaves H, Cohen KM, Wallinga J, Hoek WZ (2012) Climate-driven

fluvial development and valley abandonment at the last glacial-interglacial transition (Oude IJssel-Rhine, Germany). Neth J Geosci 91(1–2):37–62

Jochim MA, Herhahn C, Starr H (1999) The Magdalenian colonization of Southern Germany. AmAnthropol 101(1):129–142

Knappett C, Rivers R, Evans T (2011) The Theran eruption and Minoan palatial collapse: new interpre-tations gained from modelling the maritime network. Antiquity 85(329):1008–1023

Kramer D (2009) Sie haben festgestellt, dass es keinen Sommer gegeben hat. Der Ausbruch des Tambora(Indonesien) am 10. April 1815 und seine Auswirkungen. In: Schenk GJ (ed) Katastrophen. VomUntergang Pompejis bis zum Klimawandel. Jan Thorbecke Verlag, Ostfildern, pp 132–146

Kussner M (2010) The late upper Palaeolithic in the catchment area of the River Saale—facts and con-siderations. Quartar 57:125–137

Kwadijk JCJ (1991) Sensitivity of the river Rhine discharge to environmental change, a first tentativeassessment. Earth Surf Proc Land 16(7):627–637

Larsen LB, Vinther BM, Briffa KR, Melvin TM, Clausen HB, Jones PD, Siggaard-Andersen ML, HammerCU, Eronen M, Grudd H, Gunnarson BE, Hantemirov RM, Naurzbaev MM, Nicolussi K (2008) Newice core evidence for a volcanic cause of the AD 536 dust veil. Geophys Res Lett 35(4):L04708

Leesch D, Muller W, Nielsen E, Bullinger J (2012) The Magdalenian in Switzerland: re-colonization of anewly accessible landscape. Quat Int 272–273:191–208

358 Nat Hazards (2014) 71:335–362

123

Leroy SAG (2006) From natural hazard to environmental catastrophe: past and present. Quat Int 158:4–12Litt T, Schmincke H-U, Frechen M, Schluchter C (2008) Quaternary. In: McCann T (ed) The geology of

Central Europe—Mesozoic and Cenozoic, vol 2. Geological Society of London, London,pp 1287–1340

Loew S (2009) Korrespondenzanalyse und Behausungsstrukturen. Siedlungsplatzanalyse des Federmes-serfundplatzes Russelsheim 122 (Kr. Groß-Gerau, Hessen). Archaologisches Korrespondenzblatt39:309–324

Lorenz D (2013) The diversity of resilience: contributions from a social science perspective. Nat Hazards67(1):7–24

Lowe JJ, Rasmussen SO, Bjorck S, Hoek WZ, Steffensen JP, Walker MJC, Yu ZC (2008) Synchronisationof palaeoenvironmental events in the North Atlantic region during the last termination: a revisedprotocol recommended by the INTIMATE group. Quat Sci Rev 27(1–2):6–17

Lowenborg D (2012) An iron age shock doctrine—did the AD 536–537 event trigger large-scale socialchanges in the Malaren valley area? J Archaeol Anc Hist 4:1–29

Lund KA, Benediktsson K (2011) Inhabiting a risky earth: the Eyjafjallajokull eruption in 2010 and itsimpacts. Anthropol Today 27(1):6–9

Mace R, Holden CJ (2005) A phylogenetic approach to cultural evolution. Trends Ecol Evol 20(3):116–121Mann CJ (2003) Observational research methods. Research design II: cohort, cross sectional, and case-

control studies. Emerg Med J 20(1):54–60Mason BG, Pyle DM, Oppenheimer C (2004) The size and frequency of the largest explosive eruptions on

Earth. B Volcanol 66(8):735–748Mastrolorenzo G, Petrone P, Pappalardo L, Sheridan MF (2006) The Avellino 3780-year-BP catastrophe as

a worst-case scenario for a future eruption at Vesuvius. Proc Nat Acad Sci USA 103(12):4366–4370Mathiassen T (1946) En senglacial Boplads ved Bromme. Aarbøger for nordisk Oldkyndighed og Historie

1946:121–197Mauch C, Pfister C (eds) (2009) Natural disasters, cultural responses: case studies toward a global envi-

ronmental history. Lexington Books, LanhamMcCormick M, Dutton PE, Mayewski PA (2007) Volcanoes and the climate forcing of Carolingian Europe,

AD 750–950. Speculum 82:865–895McCoy FW, Heiken G (eds) (2000) Volcanic hazards and disasters in human antiquity. Geophysical

monographs no. 139. Geological Society of America, BoulderMcGuire WJ, Griffiths DR, Hancock PL, Stewart IS (eds) (2000) The archaeology of geological catastro-

phes. Geological society special publication no. 171. Geological Society, LondonMehringer PJ, Sarna-Wojcicki AM, Wollwage LK, Sheets P (2005) Age and extent of the Ilopango TBJ

tephra inferred from a Holocene chronostratigraphic reference section, Lago de Yojoa, Honduras. QuatRes 63(2):199–205

Meischner D, Gruger E (2008) Entstehung des Erdfallsees Jues in Herzberg am Harz vor 12.916 Jahren. In:Rohling H-G, Zellmer H (eds) Geo Top 2008 ‘‘Landschaften lesen lernen’’. 12. Internationale Jahr-estagung der Fachsektion GeoTop der Deutschen Gesellschaft fur Geowissenschaften, 30. April-4. Mai2008 in Konigslutter im Geopark Harz, Braunschweiger Land, Ostfalen. Schriftenreihe der DeutschenGesellschaft fur Geowissenschaften, Heft 56. Deutsche Gesellschaft fur Geowissenschaften, Hannover,p 19

Merkt J, Muller H (1999) Varve chronology and palynology of the Lateglacial in Northwest Germany fromlacustrine sediments of Hamelsee in Lower Saxony. Quat Int 61:41–59

Michel-Kerjan E (2012) How resiliant is your country? Nature 491(7425):497Mitchell P (2008) Practising archaeology at a time of climatic catastrophe. Antiquity 82(318):1093–1103Morgan MS (2013) Nature’s experiments and natural experiments in the social sciences. Philos Soc Sci

43(3):341–357Newhall CG, Self S (1982) The volcanic explosivity index (VEI): an estimate of explosive magnitude for

historical volcanism. J Geophys Res 87:1231–1238Nielsen EH (2009) Palaolithikum und Mesolithikum in der Zentralschweiz. Mensch und Umwelt zwischen

170000 und 5500 v. Chr. Archaologische Schriften Luzern 13. Kantonaler Lehrmittelverlag, LuzernNovember V (2008) Spatiality of risk. Environ Plann A 40(7):1523–1527Nowell DAG, Jones MC, Pyle DM (2006) Episodic quaternary volcanism in France and Germany. J Quat

Sci 21(6):645–675O’Keefe P, Westgate K, Wisner B (1976) Taking the naturalness out of natural disasters. Nature

260:566–567Oliver-Smith A (1996) Anthropological research on hazards and disasters. Annu Rev Anthropol 25:303–328

Nat Hazards (2014) 71:335–362 359

123

Oliver-Smith A (1999) ‘‘What is a disaster?’’ Anthropological perspectives on a persistent question. In:Oliver-Smith A, Hoffman SM (eds) The angry earth. Disaster in anthropological perspective. Routl-edge, London, pp 18–34

Oliver-Smith A (2004) Theorizing vulnerability in a globalized world: a political ecological perspective. In:Hilhorst T, Frerks G, Bankoff G (eds) Mapping vulnerability: disasters, development, and people.Earthscan Publications, London, pp 10–24

Olshansky R, Hopkins L, Johnson L (2012) Disaster and recovery: processes compressed in time. NatHazards Rev 13(3):173–178

Oppenheimer C (2003) Climatic, environmental and human consequences of the largest known historiceruption: Tambora volcano (Indonesia) 1815. Prog Phys Geog 27(2):230–259

Oppenheimer C (2011) Eruptions that shook the world. Cambridge University Press, CambridgeOrt MH, Elson MD, Anderson KC, Duffield WA, Samples TL (2008) Variable effects of cinder-cone

eruptions on prehistoric agrarian human populations in the American southwest. J Volcanol Geoth Res176(3):363–376

Pagel M, Meade A (2005) Bayesian estimation of correlated evolution across cultures: a case study ofmarriage systems and wealth transfer at marriage. In: Mace R, Holden CJ, Shennan SJ (eds) Theevolution of cultural diversity. A phylogenetic approach. UCL Press, London, pp 235–256

Park C, Schmincke H-U (1997) Lake formation and catastrophic dam burst during the late PleistoceneLaacher See Eruption (Germany). Naturwissenschaften 84:521–525

Park C, Schmincke H-U (2009) Apokalypse im rheintal. Spektrum der Wissenschaften 2009(2):78–87Pasda C (2001) Das Knochengerat vom spatpalaolithischen Fundplatz Kleinlieskow in der Niederlausitz.

Ein Essay zum steinzeitlichen Angelhaken. In: Gehlen B, Heinen M, Tillmann A (eds) Zeit-Raume.Gedenkschrift fur Wolfgang Taute. Verlag Rudolf Habelt GmbH, Bonn, pp 397–408

Pedersen R (2010) Eyjafjallajokull. Vulkanen der lammede Europa. Gyldendal, CopenhagenPettitt P (2008) The British Upper Palaeolithic. In: Pollard J (ed) Prehistoric Britain. Blackwell, Oxford,

pp 18–57Pettitt P, White MJ (2012) The British Palaeolithic. Human societies at the edge of the pleistocene world.

Routledge Archaeology of Northern Europe. Routledge, LondonPfister C (2009) Learning from nature-induced disasters: theoretical considerations and case studies from

Western Europe. In: Mauch C, Pfister C (eds) Natural disasters, cultural responses: case studies towarda global environmental history. Lexington Books, Lanham, pp 17–40

Pyle DM (2000) Sizes of volcanic eruptions. In: Sigurdsson H, Houghton BF, McNutt SR, Rymer H, Stix J(eds) Encyclopedia of Volcanoes. Academic Press, San Diego, pp 263–270

Quarantelli EL (1995) What is a disaster? Int J Mass Emerg Disasters 13(3):221–229Raynal J-P, Alborelivadie C, Piperno M (eds) (2002) Hommes et volcans. De l’eruption a l’objet. Les

dossiers de l’Archeo-Logis, Clermond-FerrandRenfrew C (1979) The eruption of Thera and Minoan Crete. In: Sheets PD, Grayson DK (eds) Volcanic

activity and human ecology. Academic Press, London, pp 565–585Riede F (2007) Der Ausbruch des Laacher See-Vulkans vor 12.920 Jahren und urgeschichtlicher Kultu-

rwandel am Ende des Allerod. Eine neue Hypothese zum Ursprung der Bromme Kultur und desPerstunien. Mitteilungen der Gesellschaft fur Urgeschichte 16:25–54

Riede F (2008) The Laacher See-eruption (12,920 BP) and material culture change at the end of the Allerødin Northern Europe. J Archaeol Sci 35(3):591–599

Riede F (2009) The loss and re-introduction of bow-and-arrow technology: a case study from the SouthernScandinavian Late Palaeolithic. Lithic Technol 34(1):27–45

Riede F (2011a) Steps towards operationalising an evolutionary archaeological definition of culture. In:Roberts BW, Vander Linden M (eds) Investigating archaeological cultures. Material culture, vari-ability, and transmission. Springer, New York, pp 245–270

Riede F (2011b) Adaptation and niche construction in human prehistory: a case study from the southernScandinavian Late Glacial. Philos Trans R Soc B 366:793–808

Riede F (2012a) A possible Brommian shaft-smoother from the site of Møllehøje, north-western Denmark.Mesolithic Miscellany 22(1):10–18

Riede F (2012b) Tephrochronologische Nachuntersuchungen am endpalaolithischen Fundplatz Rothenkir-chen, Kreis Fulda. Fuhrte der Ausbruch des Laacher See-Vulkans (10966 v. Chr.) zu einer anhaltendenSiedlungslucke in Hessen? Jahrbuch des nassauischen Vereins fur Naturkunde 133:47–68

Riede F (in press) The resettlement of northern Europe. In: Cummings V, Jordan P, Zvelebil M (eds) Oxfordhandbook of the archaeology and anthropology of hunter-gatherers. Oxford University Press, Oxford,pp xx–xx

Riede F, Bazely O (2009) Testing the ‘Laacher See hypothesis’: a health hazard perspective. J Archaeol Sci36(3):675–683

360 Nat Hazards (2014) 71:335–362

123

Riede F, Edinborough K (2012) Bayesian radiocarbon models for the cultural transition during the Allerødin southern Scandinavia. J Archaeol Sci 39(3):744–756

Riede F, Thastrup M (2013) Tephra, tephrochronology and archaeology—a (re-)view from Northern Eur-ope. Heritage Sci 1(1):15

Riede F, Wheeler JM (2009) Testing the ‘Laacher See hypothesis’: tephra as dental abrasive. J Archaeol Sci36(10):2384–2391

Riede F, Bazely O, Newton AJ, Lane CS (2011) A Laacher See-eruption supplement to Tephrabase:investigating distal tephra fallout dynamics. Quat Int 246(1–2):134–144

Ritter JRR, Jordan M, Christensen UR, Achauer U (2001) A mantle plume below the Eifel volcanic fields,Germany. Earth Planet Sci Lett 186:7–14

Roberts BW, Vander Linden M (eds) (2011) Investigating archaeological cultures. Material culture, vari-ability, and transmission. Springer, New York

Rolett B, Diamond J (2004) Environmental predictors of pre-European deforestation on Pacific islands.Nature 431(7007):443–446

Scarre C (ed) (2005) The human past. World prehistory and the development of human societies. Thames &Hudson, London

Schenk GJ (ed) (2009) Katastrophen. Vom Untergang Pompejis bis zum Klimawandel. Jan ThorbeckeVerlag, Ostfildern

Schmincke H-U (2006) Environmental impacts of the Lateglacial eruption of the Laacher See Volcano,12.900 cal BP. In: von Koenigswald W, Litt T (eds) 150 years of Neanderthal discoveries, vol 2. TerraNostra, Bonn, pp 149–153

Schmincke H-U, Hinzen K-G (2008) Vulkanismus und Erdbeben. In: Felgentreff C, Glade T (eds) Natur-risiken und Sozialkatastrophen. Spektrum Akademischer Verlag, Berlin, pp 141–150

Schmincke H-U, Park C, Harms E (1999) Evolution and environmental impacts of the eruption of LaacherSee Volcano (Germany) 12,900 a BP. Quat Int 61:61–72

Schreiber UC (2006) Die Flucht der Ameisen. Piper, Munchen/ZurichSchwendler RH (2012) Diversity in social organization across Magdalenian Western Europe ca. 17–12,000

BP. Quat Int 272–273:333–353Self S (2006) The effects and consequences of very large explosive volcanic eruptions. Philos Trans R Soc

A 364(1845):2073–2097Sheets PD (1979) Maya recovery from volcanic disasters—Ilopango and Ceren. Archaeology 32(3):32–42Sheets PD (1981) Volcanoes and the Maya. Nat Hist 90(8):32–41Sheets PD (1999) The effects of explosive volcanism on ancient egalitarian, ranked, and stratified societies

in Middle America. In: Oliver-Smith A, Hoffman SM (eds) The angry earth. Disaster in anthropo-logical perspective. Routledge, London, pp 36–58

Sheets PD (2001) The effects of explosive volcanism on simple to complex societies in ancient MiddleAmerica. In: Markgraf V (ed) Interhemispheric climate linkages. Academic Press, London, pp 73–86

Sheets PD (2007) People and Volcanoes in the Zapotitan Valley, El Slavador. In: Grattan J, Torrence R (eds)Living under the shadow. Cultural impacts of volcanic eruptions. Left Coast Press, Walnut Creek,pp 67–89

Sheets PD (2008) Armageddon to the Garden of Eden: explosive volcanic eruptions and societal resiliencein ancient Middle America. In: Sandweiss D, Quilter J (eds) El Nino, catastrophism, and culturechange in ancient America. Harvard University Press, Cambridge, pp 167–186

Sheets PD (2012) Responses to explosive volcanic eruptions by small to complex societies in ancientMexico and Central America. In: Cooper J, Sheets PD (eds) Surviving sudden environmental change.University of Colorado Press, Boulder, pp 43–63

Small C, Naumann T (2001) The global distribution of human population and recent volcanism. EnvironHazards 3(3/4):93–109

Sørensen L (2010) The Laacher See volcanic eruption. Challenging the idea of cultural disruption. ActaArchaeol-Den 81(1):270–281

Stapert D (2000) The Late Palaeolithic in the northern Netherlands. In: Valentin B, Bodu P, Christensen M(eds) L’Europe centrale et septentrionale au Tardiglaciaire. Actes de la Table-Ronde internationale deNemours, 13–16 mai 1997. Memoires du Musee de Prehistoire d’Ile de France, Paris, pp 175–195

Stommel H, Stommel E (1983) Volcano weather. The story of 1816, the year without a summer. Seven SeasPress, New Port

Street M, Joris O, Turner E (2012) Magdalenian settlement in the German Rhineland—an update. Quat Int272–273:231–250

Stump D (2013) On Applied Archaeology, Indigenous Knowledge, and the Usable Past. Curr Anthropol54(3):268–298

Nat Hazards (2014) 71:335–362 361

123

Textor C, Sachs PM, Graf H-F, Hansteen TH (2003) The 12 900 years BP Laacher See eruption: estimationof volatile yields and simulation of their fate in the plume. Geol Soc Spec Publ 213(1):307–328

Toft B, Reynolds S (1994) Learning from disasters. A management approach. Butterworth-Heinemann,Oxford

Torrence R, Grattan J (eds) (2002) Natural disasters and cultural change. Routledge, LondonVan de Noort R (2011) Conceptualising climate change archaeology. Antiquity 85(329):1039–1048van den Bogaard P (1995) 40Ar/39Ar ages of sanidine phenocrysts from Laacher See Tephra (12,900 year

BP): chronostratigraphic and petrological significance. Earth Planet Sci Lett 133(1–2):163–174van den Bogaard P, Schmincke H-U (1984) The eruptive center of the late quaternary Laacher See Tephra.

Geol Rundsch 73(3):933–980van den Bogaard P, Schmincke H-U (1985) Laacher See Tephra: a widespread isochronous late quaternary

tephra layer in Central and Northern Europe. Geol Soc Am Bull 96(12):1554–1571van den Bogaard P, Schmincke H-U, Freundt A, Park C (1990) Evolution of complex plinian eruptions: the

Late Quaternary Laacher See case history. In: Hardy DA, Keller J, Galanopoulos VP, Flemming NC,Druitt TH (eds) Thera and the Aegean world III. Earth Sciences, vol 2. The Thera Foundation, London,pp 463–483

van Raden UJ, Colombaroli D, Gilli A, Schwander J, Bernasconi SM, van Leeuwen J, Leuenberger M,Eicher U (in press) High-resolution late-glacial chronology for the Gerzensee lake record (Switzer-land): d18O correlation between a Gerzensee-stack and NGRIP. Palaeogeography, Palaeoclimatology,Palaeoecology

Veil S, Breest K, Grootes PM, Nadeau M-J, Huls M (2012) A 14 000-year-old amber elk and the origins ofnorthern European art. Antiquity 86(333):660–673

Vitaliano DB (2007) Geomythology: geological origins of myths and legends. Geol Soc Spec Publ273(1):1–7

Warburton D, Heinemeier J, Friedrich WL (eds) (2009) Time’s up! Dating the Minoan eruption of Santorini.Acts of the Minoan Eruption Chronology Workshop, Sandbjerg November 2007. Monographs of theDanish Institute at Athens, vol 10. Aarhus University Press, Arhus

Weniger G-C (1989) The Magdalenian in Western Central Europe: settlement pattern and regionality.J World Prehist 3(3):323–372

White GF (1974) Natural hazards research: concepts, methods and policy implications. In: White GF (ed)Natural hazards: local, national, global. Oxford University Press, Oxford, pp 3–16

Wisner B (2010) Climate change and cultural diversity. Int Soc Sci J 61(199):131–140Wisner B, Blaikie P, Cannon T, Davis I (2004) At risk. Natural hazards, people’s vulnerability and disasters,

2nd edn. Routledge, LondonWitham CS (2005) Volcanic disasters and incidents: a new database. J Volcanol Geoth Res

148(3–4):191–233Worner G, Schmincke H-U (1984a) Mineralogical and Chemical Zonation of the Laacher See Tephra

Sequence (East Eifel, W. Germany). J Petrol 25(4):805–835Worner G, Schmincke H-U (1984b) Petrogenesis of the Zoned Laacher See Tephra. J Petrol 25(4):836–851Yoon D (2012) Assessment of social vulnerability to natural disasters: a comparative study. Nat Hazards

63(2):823–843Zhu H, Bozdag E, Peter D, Tromp J (2012) Structure of the European upper mantle revealed by adjoint

tomography. Nat Geosci 5(7):493–498

362 Nat Hazards (2014) 71:335–362

123