You are on page 1of 342

Issues and Concepts in Historical Ecology

Historical ecology is a research framework which draws upon


diverse evidence to trace complex, long-term relationships between
humanity and Earth. With roots in anthropology, archaeology,
ecology and palaeoecology, geography, and landscape and heritage
management, historical ecology applies a practical and holistic
perspective to the study of change. Furthermore, it plays an
important role in both fundamental research and in developing
future strategies for integrated, equitable landscape management.
The framework presented in this volume covers critical issues,
including: practicing transdisciplinarity, the need for understanding
interactions between human societies and ecosystem processes,
the future of regions and the role of history and memory in a
changing world. Including many examples of co-developed research,
Issues and Concepts in Historical Ecology provides a platform
for collaboration across disciplines and aims to equip researchers,
policy-makers, funders, and communities to make decisions that
can help to construct an inclusive and resilient future for humanity.

Carole L. Crumley is Emerita Professor of Anthropology at the


University of North Carolina, Chapel Hill, and Senior Researcher at
the Swedish Biodiversity Centre, Uppsala, Sweden. She is a founder
of historical ecology, editing Historical Ecology: Cultural Knowledge
and Changing Landscapes, 1994. Her research interests are broad
in both science and the humanities including: anthropology,
archaeology, landscape ecology, palaeoecology and climatology.

Tommy Lennartsson is Biologist and an Associate Professor in


Conservation Biology at the Swedish University of Agricultural
Sciences, Uppsala, and Researcher at the Swedish Biodiversity
Centre. His research focuses on applied aspects of biodiversity
conservation and use of natural resources, with a historical
perspective and in relation to climate change.

Anna Westin is Biologist and an Associate Professor in Agricultural


History at the Swedish University of Agricultural Sciences, and
Researcher at the Swedish Biodiversity Centre. Her research uses
historical ecology to study biodiversity, ecology, cultural heritage
and history.

Downloaded from https://www.cambridge.org/core. The University of British Columbia Library, on 02 Aug 2018 at 17:08:21, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
ii

Downloaded from https://www.cambridge.org/core. The University of British Columbia Library, on 02 Aug 2018 at 17:08:21, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
i

Issues and Concepts


in Historical Ecology
The Past and Future of
Landscapes and Regions

Edited by
Carole L. Crumley
University of North Carolina, Chapel Hill, NC,
USA; and Swedish Biodiversity Centre, Uppsala, Sweden

Tommy Lennartsson
Swedish Biodiversity Centre, Uppsala, Sweden

Anna Westin
Swedish Biodiversity Centre, Uppsala, Sweden

Downloaded from https://www.cambridge.org/core. The University of British Columbia Library, on 02 Aug 2018 at 17:08:21, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
v
i

University Printing House, Cambridge CB2 8BS, United Kingdom

One Liberty Plaza, 20th Floor, New York, NY 10006, USA

477 Williamstown Road, Port Melbourne, VIC 3207, Australia

4843/24, 2nd Floor, Ansari Road, Daryaganj, Delhi – 110002, India

79 Anson Road, #06-04/06, Singapore 079906

Cambridge University Press is part of the University of Cambridge.

It furthers the University’s mission by disseminating knowledge in the pursuit of


education, learning, and research at the highest international levels of excellence.

www.cambridge.org
Information on this title: www.cambridge.org/9781108420983
DOI: 10.1017/9781108355780

© Cambridge University Press 2018

This publication is in copyright. Subject to statutory exception


and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without the written
permission of Cambridge University Press.

First published 2018

Printed in the United Kingdom by TJ International Ltd. Padstow Cornwall

A catalogue record for this publication is available from the British Library.

ISBN 978-1-108-42098-3 Hardback

Cambridge University Press has no responsibility for the persistence or accuracy


of URLs for external or third-party internet websites referred to in this publication
and does not guarantee that any content on such websites is, or will remain,
accurate or appropriate.

Downloaded from https://www.cambridge.org/core. The University of British Columbia Library, on 02 Aug 2018 at 17:08:21, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
v

Contents

List of Illustrations page vii


List of Contributors xiii

1 Is There a Future for the Past? 1


Carole L. Crumley, Anna Westin, and
Tommy Lennartsson

PART I CHALLENGES: TIME AND MEMORY

2 Historical Ecology and the Longue Durée 13


Paul Sinclair, Jon Moen, and
Carole L. Crumley

3 Human and Societal Dimensions of Past


Climate Change 41
Fredrik Charpentier Ljungqvist

4 Rural Communities and Traditional


Ecological Knowledge 84
Anamaria Iuga, Anna Westin, Bogdan Iancu,
Monica Stroe, and Håkan Tunón

5 Baselines and the Shifting Baseline Syndrome –


Exploring Frames of Reference in Nature
Conservation 112
Tuija Hilding- Rydevik, Jon Moen, and
Carina Green

Downloaded from https://www.cambridge.org/core. The University of British Columbia Library, on 02 Aug 2018 at 17:17:42, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
vi

vi Contents

PART II APPROACHES: CONCEPTS AND METHODS

6 Concepts for Integrated Research in Historical Ecology 145


Ove Eriksson, Anneli Ekblom, Paul Lane,
Tommy Lennartsson, and Karl- Johan
Lindholm

7 Diversity in Ecological and Social Contexts 182


Tommy Lennartsson, Ove Eriksson,
Anamaria Iuga, Jesper Larsson, Jon Moen,
Michael D. Scholl, Anna Westin, and
Carole L. Crumley

8 How to Operationalise Collaborative Research 240


Elizabeth A. Jones, Anna Westin,
Scott Madry, Seth Murray, Jon Moen, and
Amanda Tickner

PART III MOVING FORWARD

9 Historical Ecology in Theory and Practice:


Editors’ Reflections 275
Tommy Lennartsson, Anna Westin, and
Carole L. Crumley

10 Taking Research into Action in Historical Ecology 298


Carole L. Crumley

Index 315

Colour plates are to be found between pp. 144 and 145.

Downloaded from https://www.cambridge.org/core. The University of British Columbia Library, on 02 Aug 2018 at 17:17:42, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
vi

Illustrations

1.1 Contributors’ major fields of study page 4


2.1 A wedge of starlings 27
2.2 Time as a braided river with different processes
flowing at different rates 33
3.1 Main characteristics of the climate variability in the
Northern Hemisphere from the end of the last glacial
period until present times 44
3.2 Reconstructed temperature variability in Europe and
China, respectively, during the past two millennia 50
3.3 Reconstructed long-term aridity changes (AD 800–
2003) in the western half of the United States derived
from the tree-ring width–based North American
Drought Atlas 52
3.4 Reconstructed relative strength of the summer
monsoon in central China and northeast India,
respectively, during the past one to two millennia
from stable isotope cave speleothem data 53
3.5 Schematic overview of causal linkages between
climate changes with adverse changes for agricultural
productivity and human/societal crises in the
premodern world 59
3.6 The different orders of impact from adverse climate
change or adverse extreme weather events ranging
from the direct biophysical effects on plant growth
down to cultural responses 62
4.1 The building of a haystack in Botiza (Romania), 2014 86
4.2 Sharpening the scythe during mowing. Botiza
(Romania), 2010 87

vii

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:20:32, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
vi

viii Illustrations

4.3 During the milk-measurement custom in Șurdești


village (Romania), the milk from all animals is
measured in order to divide input and output during
the season, 2009 93
4.4 Collective work of mowing. Fundata (Romania), 2014 94
4.5 The family of Lars Olsson with extra hands is
mowing the meadows of the Småland farm, county of
Jämtland, Sweden, 31 July 1916 95
4.6 ‘Petrified haystacks’ in Șurdești village (Romania).
The legend tells that the haystacks were turned into
stone by St. Peter, who encountered a man working
with hay on a forbidden day 100
4.7 The custom celebrates the first man who ploughed
his land, in Șurdești village (Romania). After it was
forbidden during the Communist period, the custom
was revitalised in the 1990s. The photo presents the
moment when the celebrated man is taken to the river
and his head bathed in water, in the belief that this
would have a good influence on the crops. Șurdești
(Romania), 2012 106
4.8 A traditional wooden fence gärdesgård is under
construction in the archipelago of Stockholm. The
knowledge of how to build the gärdesgård was
widespread since the duty and interest to maintain
them was divided among all farmers in the village.
Today fairly few people master these skills in Sweden,
and most fences are of different and more modern
types. Hjälmö farm (Sweden), 2014 107
5.1 Photo of a Sami family, probably from the late
nineteenth century. The picture is taken in the central
part of what is now the Tjeggelvas nature reserve and
is currently considered an example of a pristine boreal
forest, i.e., untouched by humans 119

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:20:32, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
x

Illustrations ix

5.2 Ideas and memories are shaped through many different


sources, e.g., art and paintings portraying a pristine
nature. Artwork by Jan Brueghel the Younger – 1601–78 132
5.3 Memories of extreme catches can influence the
perception of trends or earlier states. Artwork by
Pieter Brueghel the Elder – 1526/1530–69 133
6.1 Aerial view of occupied and abandoned pastoralist
bomas in Amboseli (Kenya), January 2009, showing
visual differences in their higher nutrient content
relative to background soils 157
6.2 Dung-enriched soil in a recently abandoned pastoralist
boma, Amboseli (Kenya), January 2009 158
6.3 Managed trees (Acacia erioloba) at a pastoral well
site (Ozombu Zo Vindimba) in the Kalahari, eastern
Namibia. The tree has new and several older marks
indicating that branches have been cut. Reasons for
cutting branches are to harvest browse, firewood
or timber from trees without killing them, aiming
for the full re-growth of the tree. The local herders
explain the trees’ curved shape as a result of when the
trees were younger they were often been bent down
towards the ground for feeding calves with the highly
nutritious seed pods 159
6.4 Aerial photo from the province of Uppland, Sweden,
showing a mosaic of nature types formed by various
combinations of human impact and natural processes.
In the front, grazing has shaped a fodder-producing
grassland out of a seashore reed belt, in which the
biomass-producing vegetation is influenced by both
the grazing and the water 163
6.5 Lybby village in the province of Närke, Sweden,
showing a pasture with pollarded and formerly
pollarded trees 164

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:20:32, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
x

x Illustrations

6.6 Remnants of stone walls used for fencing out


livestock from crop fields and hay meadows (from the
province of Halland, Sweden) 164
7.1 Reindeer summer pastures at Bietsávrre settlement,
Lappland, Sweden 194
7.2 A reindeer herd on winter pastures in the boreal
forest, Sweden 195
7.3 An Indian summer view of the farm and stock of
James C. Cornell, 1848 200
7.4 Comparison of acreage of wheat, meslin (mixed wheat
and rye), and rye, Chester County, USA 202
7.5 A Swedish summer farm community consisted of
members from many households using the same
settlement and area. Herders, cows, and goats at
Björnbergets summer farm in the Parish of Leksand,
Sweden, early 20th Century 209
7.6 Detailed rules for where and when the grazing cows
could be herded were necessary institutions for
a sustainable use of woodland and mire pastures.
Female herders following the herd from Björnbergets
summer farm in the parish of Leksand, Sweden, early
twentieth century 210
7.7 A grassland landscape, stretching from garden
plantations to fields, meadows, and hilltop pastures,
showing a great diversity of grassland ecosystems 216
7.8 The use of the Carpathian agricultural landscape
in Romania is based on mixed pastoralism. The
diversity of pastures and meadows provides feed for
grazing livestock, which produce cheese and other
dairy products as well as manure for the arable fields.
The grazers have a key function in the system by
transforming grass into human food. As a by-product,
rich biological diversity is formed. Rodna Mountains,
Romania 217

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:20:32, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
xi

Illustrations xi

8.1 Farmers, researchers, and government officials


meet in the field to discuss the history and future
management of the landscape. Hjälmö farm, county of
Stockholm, Sweden 244
8.2 Researchers studying a farm archive in the house
where the documents were written, together with
descendants of the document author. Hyttbäcken
farm, county of Dalarna, Sweden 245
8.3 The logotype for the HagmarksMISTRA research
programme, reflecting different disciplines and
perspectives used 247
8.4 Maps of mill and pond area at Bauzot, France, 1838,
1912, 1983, 2002 261
8.5 Postcard of Bauzot Mill building and pond circa 1906 262
8.6 Schematic of issues to be formally addressed in
collaborative research 266
9.1 Giant pollarded beech trees in a dense forest reveal the
land-use history as well as the former forest structure.
This forest in Botiza in the Romanian Carpathians
has probably been a pasture with scattered pollards,
harvested for leaf fodder 285
9.2 In the pre-industrial Swedish landscape, it was
common that midfield islets, used for grazing, were
fenced together with the arable fields. These islets –
low hills – could not be grazed until after harvest
(August–September), and that grazing regime favoured
plant species that flower early and set seeds before
grazing. One example is the cow slip, Primula veris,
whose presence indicates the historical timing of
grazing, long after management has changed. Hjälmö,
Stockholm archipelago, Sweden 286

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:20:32, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
xi

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:20:32, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
xi

Contributors

Fredrik Charpentier Ljungqvist, Department of History and Bolin


Centre for Climate Research, Stockholm University, Stockholm,
Sweden; and Scientific Steering Committee, Integrated History
and Future of People on Earth (IHOPE), Uppsala University,
Uppsala, Sweden

Carole L. Crumley, Senior Researcher, Swedish Biodiversity Centre,


Swedish University of Agricultural Sciences (SLU) and Uppsala
University, Uppsala, Sweden; Executive Director, Integrated History
and Future of People on Earth (IHOPE), Department of Archaeology
and Ancient History, Uppsala University, Uppsala, Sweden; and
Department of Anthropology, University of North Carolina, Chapel
Hill, NC, USA

Anneli Ekblom, Department of Archaeology and Ancient History,


Uppsala University, Uppsala, Sweden; and Scientific Steering
Committee, Integrated History and Future of People on Earth
(IHOPE), Uppsala University, Uppsala, Sweden

Ove Eriksson, Department of Ecology, Environment and Plant


Sciences, Stockholm University, Stockholm, Sweden

Carina Green, Swedish Biodiversity Centre, Swedish University


of Agricultural Sciences (SLU) and Uppsala University; and
Department of Cultural Anthropology and Ethnology, Uppsala
University, Uppsala, Sweden

Tuija Hilding-Rydevik, Swedish Biodiversity Centre at the Swedish


University of Agricultural Sciences (SLU) and Uppsala University,
Uppsala, Sweden
xiii

Downloaded from https://www.cambridge.org/core. University of Exeter, on 02 Aug 2018 at 17:30:14, subject to the Cambridge Core terms of
use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
v
xi

xiv Contributors

Bogdan Iancu, Department of Sociology, National School of


Political Studies and Public Administration, Bucharest, Romania;
and National Museum of the Romanian Peasant, Bucharest,
Romania

Anamaria Iuga, Ethnology Studies Department, National Museum


of the Romanian Peasant, Bucharest, Romania

Elizabeth A. Jones, Research Laboratories of Archaeology,


University of North Carolina, Chapel Hill, NC, USA

Paul Lane, Department of Archaeology and Ancient History,


Uppsala University, Uppsala, Sweden; and School of Geography,
Archaeology and Environmental Studies, University of the
Witwatersrand, Johannesburg, South Africa

Jesper Larsson, Royal Swedish Academy of Letters, History and


Antiquities; Swedish University of Agricultural Sciences (SLU),
Uppsala, Sweden; and Affiliated Faculty, Ostrom Workshop, Indiana
University, Bloomington, IN, USA

Tommy Lennartsson, Swedish Biodiversity Centre at the Swedish


University of Agricultural Sciences (SLU) and Uppsala University,
Uppsala, Sweden

Karl-Johan Lindholm, Department of Archaeology and Ancient


History, Uppsala University, Uppsala, Sweden

Scott Madry, Department of Anthropology, University of North


Carolina, Chapel Hill, NC, USA

Jon Moen, Department of Ecology and Environmental Science,


Umeå University, Umeå, Sweden

Downloaded from https://www.cambridge.org/core. University of Exeter, on 02 Aug 2018 at 17:30:14, subject to the Cambridge Core terms of
use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
v
x

Contributors xv

Seth Murray, Department of Sociology and Anthropology, North


Carolina State University, Raleigh, NC, USA

Michael D. Scholl, Department of Anthropology, Wagner College,


Staten Island, New York, NY, USA; and Department of Sociology
and Anthropology, College of Staten Island, City University of
New York, New York, NY, USA

Paul Sinclair, Department of Archaeology and Ancient History,


Uppsala University, Uppsala, Sweden; and Department of
Anthropology and Archaeology, Pretoria University, Pretoria, South
Africa; and Executive Committee, Integrated History and Future of
People on Earth (IHOPE), Uppsala University, Uppsala, Sweden

Monica Stroe, Department of Sociology, National University of


Political Studies and Public Administration, Bucharest, Romania

Amanda Tickner, Map Library, MSU Libraries, Michigan State


University, MI, USA

Håkan Tunón, Swedish Biodiversity Centre at the Swedish


University of Agricultural Sciences (SLU) and Uppsala University,
Uppsala, Sweden; and Royal Swedish Academy of Forestry and
Agriculture, Stockholm, Sweden

Anna Westin, Swedish Biodiversity Centre at the Swedish


University of Agricultural Sciences (SLU) and Uppsala University,
Uppsala, Sweden

Downloaded from https://www.cambridge.org/core. University of Exeter, on 02 Aug 2018 at 17:30:14, subject to the Cambridge Core terms of
use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
xvi

Downloaded from https://www.cambridge.org/core. University of Exeter, on 02 Aug 2018 at 17:30:14, subject to the Cambridge Core terms of
use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
1

1 Is There a Future for the Past?


Carole L. Crumley, Anna Westin, and
Tommy Lennartsson

1.1 Introduction
The health of our planet and our species is jeopardised by a host of
dangers to environmental, political, and social security brought on
by a now-familiar litany: anthropogenic pollution, climate change,
increasing inequity and conflict, and the global-scale loss of biodi-
versity. These circumstances require attention both locally, where
global changes impact communities, and globally, where local prac-
tices influence global change. We humans are only part of a complex
network of elements and relations that make up planet Earth. Within
this enormous and essentially closed ecosystem, our lives are influ-
enced by events, processes, and conditions that began long before the
first humans and that will outlast humanity.
Now we have entered the Anthropocene, an era when human
activity must be considered a major component (a ‘driver’) of global
environmental change. The dynamic, nonlinear system in which we
live is not in equilibrium and does not act in a predictable manner.
If our own and other species are to continue to thrive, it is of utmost
importance that we identify the conditions, ideas, and practices that
nurture both the planet and the species that live on it. Our best lab-
oratory for this is the past, where long-, medium-, and short-term
variables can be identified and their roles evaluated. Perhaps the past
is our only laboratory: experimentation requires time we no longer
have. Thus the integration of our understanding of human history
with that of the Earth system is a timely and urgent task.
Historical ecology traces the complex relationships between
our species and Earth, examined over the long term. Its roots are in
a variety of disciplines (e.g. anthropology, archaeology, ecology and

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:47:46, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.001
2

2 Crumley et al.

palaeoecology, geography, landscape and heritage management, for-


est history). Practitioners take the term ecology to include humans
as a component of all ecosystems, and the term history to include
that of the Earth system as well as the social and physical past of our
species.
Historical ecology is neither a discipline nor a theory. It is a
research framework for merging many kinds of evidence (e.g. doc-
uments, archaeology, ethnography, ecology, and a broad array of
environmental studies) to reach new understandings about the
human-environment relationship. Research is stimulated by new
questions, insights, and methods for combining and analysing dif-
ferent sources of information. Historical ecology is an emerging
field of study that can facilitate interdisciplinarity and generate new
understandings.
Historical ecology draws on a broad spectrum of theories, con-
cepts, issues, and evidence from the biological and physical sciences,
ecology, the social sciences and humanities, and from diverse commu-
nities of practice. Historical ecologists are not concerned with battles
among various theories: rather, they take a holistic, practical, and
dialectical perspective both on the study of change and on the prac-
tice of interdisciplinary research. This framework enables flexible,
multiple, and inclusive narratives of human-environment relations
over time in a particular geographic location (hence an emphasis on
landscape). The goal of historical ecology is to use scholarly and other
varieties of knowledge in concert, so that management decisions that
will shape the human future on Earth can be effective and equitable.

1.2 Integrating Knowledges, Exploring


Concepts and Issues
In Greek mythology Ariadne, daughter of King Minos, gave Theseus
a red thread, which he unwound as he entered the sacred labyrinth
of the Minotaur and thereby found his way through its complexities.
We who grapple with the complexities of connecting the human
and biophysical past – much less using our insights to chart the

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:47:46, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.001
3

Is there a Future for the Past? 3

future – can use a few red threads, concepts, and premises to guide
our work. But while combining different kinds of knowledge about
the world around us is an admirable goal in the abstract, many ques-
tions, issues, and concepts remain that must first be addressed to the
satisfaction of collaborators.
Our work on this book has been sponsored by the Centre for
Biodiversity (CBM), begun a quarter century ago as a joint venture
between the Swedish University of Agricultural Sciences (SLU) and
neighbouring Uppsala University. To meet the requirements of the
Convention on Biological Diversity, the Swedish Parliament tasked
CBM to create an innovative environmental management and pol-
icy structure for Sweden. CBM now works closely with government
agencies, policy-makers, natural resource managers, local and indig-
enous communities, and other stakeholders in Sweden, Europe, and
elsewhere in the world.
In keeping with CBM’s mandate, the editors of this book
(Crumley, Lennartsson, Westin) convened a three-day workshop in
2013 on historical ecology, one of the pillars of CBM practice. The
participants are from Great Britain, Romania, Sweden, and the United
States; about a third of us have ties to CBM. Participants were cho-
sen for their experience with the historical ecology framework and
their ‘cross-training’ through fieldwork and in areas of study beyond
their own core training (see Figure 1.1). We are twenty-two historical
ecologists who chose both the subject matter and the structure of
this book, which we have written together.
Contributors to this book have wide-ranging research expe-
rience, which has been accrued all over the world. Many work in
Europe, including Scandinavia (Greenland, Iceland, Norway, Sweden),
Western Europe (France, Great Britain, Greece, Ireland, Italy, Scotland,
Spain), and Eastern Europe (Romania, the Danube basin countries);
many contributors work on issues related to the European Union.
The African continent is well represented: Botswana, the Comores,
Egypt, Kenya, Madagascar, Mozambique, Namibia, Somalia, South
Africa, Tanzania, Zimbabwe, and the Sahel. Many contributors

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:47:46, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.001
4

4 Crumley et al.

Figure 1.1 Contributors’ major fields of study.

have Western Hemisphere experience (Canada, the United States,


the Rockies and the Great Lakes, Bolivia). South East Asia (Laos,
the Maldives, Sri Lanka, Thailand), Australia, New Zealand, and
Antarctica are also represented.
Using an historical ecology analytic framework is never easy
and problem-free; there are always difficulties, such as in merging
data sets (such as quantitative and qualitative information), finding
proxies for certain data (such as using parish records for population),
and many others.
Before the workshop, the editors circulated some questions
that are basic to research in historical ecology. While this book does
not address all of the questions directly, we hope to have contributed
to their discussion.

• How can research questions and approaches be combined to generate new


knowledge?
• How can mutually acceptable research questions be framed?
• What information is missing from a research design? How can it
be found?
• How should one ask questions about data and sources of colleagues in
other disciplines?

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:47:46, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.001
5

Is there a Future for the Past? 5

• How can the historical ecology approach be adapted to specific questions,


times, and places?
• How can historical ecology address multiple spatial scales (e.g. micro-to-
macro economy, species-to-landscapes) and temporal scales (e.g. annual to
long-term changes)?
• How to assess the added value of merged knowledge sources (e.g.
scholarly, applied, ethical)?
• What areas of applied historical ecology are particularly fruitful or
problematic?

The editors then asked participants to share a ‘puzzle’ – a problem


or question they had or still have – of particular importance to that
person’s research; these were circulated to the group. We listened to
short presentations of each person’s puzzle. Workshop participants
sorted the puzzles into themes, which focussed the discussion; we
then constructed the structure of this volume.
We chose three themes under which we could offer multiple-
authored chapters on problematic areas in the practice of histori-
cal ecology. The long-term history of biocultural diversity would
address key issues in the study of linked human-environment rela-
tions, including how long-term studies can yield novel insights and
what the future may hold for regions and communities. Constructive
approaches would address the language and logic of troublesome
concepts and take a pragmatic approach to collaboration and data
integration. Dialogues: communities of purpose would explore some
actual communities and their inhabitants, along with how historical
ecology could be better situated in policy and governance, and the
role of emotional and spiritual connections with the natural world.
While this book’s structure has evolved, these key themes or ‘red
threads’ have guided the chapters’ evolution in the hands of multiple
authors.
The workshop only began our collaboration; there have been
five more workshops for the entire group and many more intimate
meetings of co-authors. Subsequent workshops (with participants
attending in person or electronically) allowed discussion of chapter

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:47:46, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.001
6

6 Crumley et al.

drafts between the chapters’ authors and other members of the group;
these meetings also revisited the structure of this volume and other
business. This book’s chapters are now grouped as Challenges:
Time and Memory (Chapters 2 to 5), followed by Approaches:
Concepts and Methods (Chapters 6 to 8), and Moving Forward
(Chapters 9 and 10).
Challenges addresses some essential, but usually neglected
issues related to time. Chapter 2, Historical ecology and the longue
durée, introduces the study of time, integral to disciplines that lay
claim to an understanding of society-environment dynamics. The
authors review the history of perceptions of time in time-sensitive
disciplines, then examine shortcomings of the traditional linear
approach, and finally engage with the relatively new vision of time
that has emerged with the study of complex systems. They do not
abandon linear time, but instead advocate an increased awareness of
the multiplicity of times and the exploration of their implications for
learning from the past. Chapter 3, Human and societal dimensions of
past climate change, introduces another central concept, the import-
ance of understanding key human-environment interactions over
time and space through the lens of climate and weather. Focussing
on how the impact of climate change has been met, rather than on
climate change itself, the chapter demonstrates the importance of
multiple strategies, including social responses, for enhancing resili-
ence in the face of climate change and other environmental impacts.
Chapter 4, Rural communities and traditional ecological know-
ledge, explores the emergence, transformation, and disappearance of
traditional ecological knowledge (TEK) in relation to the marking of
time in agricultural rural communities. Examples of current and past
TEK from Romania and Sweden are used to illustrate how TEK in
one place can provide insight for historical ecology elsewhere, the
implications of change and loss of TEK, and its current re-evaluation.
Chapter 5, Baselines and the shifting baseline syndrome – explor-
ing frames of reference in nature conservation, focuses on factors
that shape the quite different goals of environmental management
and environmental conservation policy, and on the role of human

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:47:46, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.001
7

Is there a Future for the Past? 7

cognition when setting baselines in nature conservation. Since base-


lines are socially constructed, norms and perceptions of the environ-
ment can change across generations, leading to the ‘shifting baseline
syndrome’ (SBS). Baselines and SBS, developed primarily for use in
the natural sciences, can benefit from memory research in the social
sciences and from the broader scope of historical ecology.
Approaches puts concepts and methods to work in the practice
of historical ecology. Chapter 6, Concepts for integrated research in
historical ecology, explores new frameworks for research developed
through a reappraisal of influential concepts in landscape ecology and
archaeology, such as adaptation, niche construction, domestication,
and entanglement. Perspectives from different disciplines on those
concepts are compared and blended, treating landscapes themselves
as agents, entangled between human and non-human, biological and
non-biological agents and processes. Based on examples from Africa
and Europe, the authors argue that careful borrowing of concepts can
lead to fruitful research in both landscape ecology and social history,
and to new conceptual frames for collaboration. Chapter 7, Diversity
in ecological and social contexts, addresses the variation in human
societies (cultural diversity) and in ecosystems (biological diversity).
Cultural and biological diversity show differences and similarities
regarding how diversity is differently framed, formed, and maintained
in natural and social sciences and in the humanities. New insights
can be achieved by comparing and merging the perspectives; an
example is the intersection between cultural and biological diversity,
termed biocultural diversity. Based on four examples of land-use sys-
tems, the authors discuss how understanding the reasons for the for-
mation or loss of biocultural diversity can help us interpret the past
and plan for the future. Chapter 8, How to operationalise collabora-
tive research, argues that, by definition, research in historical ecol-
ogy requires an interdisciplinary theoretical framework, and in most
cases a collaborative research approach as well. The conceptual and
practical aspects of interdisciplinary collaboration and integration
can be daunting, and most researchers are not trained to work in this
way. The authors provide practical considerations and suggestions

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:47:46, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.001
8

8 Crumley et al.

for collaborative interdisciplinary work, based on examples drawn


from their research experiences in historical ecology.
Moving Forward addresses the role of historical ecology in
fashioning an environmentally sound and socially supportive future.
In Chapter 9, Historical ecology in theory and practice: editors’
reflections, the editors consider ideas from the chapters, relation-
ships among the chapters, and historical ecology in a wider context.
Among the issues addressed are how historical ecology can be used
in biodiversity and cultural heritage conservation and as a guide for
sustainable resources use; how diverse approaches to time might be
reconciled; and how the engaged agenda of historical ecology can find
support among practitioners. Chapter 10, Taking research into action
in historical ecology, suggests that study of the past can expose how
the fabric of human societies and their environments is woven.
The failure of international policy to make significant progress in
slowing the drivers of climate change has necessitated a grass-roots
response: broad networks are forming and integrative research is well
under way. The emergent, collaborative, transdisciplinary, and trans-
temporal research framework of historical ecology can provide an
arena for addressing the critical issues facing humanity.
Of course issues have arisen that challenged participants to
struggle to understand one another. One participant was uncomfort-
able with the word system, common in the environmental sciences,
but problematic for some social scientists and humanists. The need
to clarify the dynamic among bio-geo-physical and human actors led
to many conversations about agency. Sometimes scholarly styles,
reflecting variation in national, disciplinary, or linguistic traditions,
required careful navigation. In particular, the divergent histories of
relevant research and collaboration in Europe and North America
prompted interesting and ultimately productive adjustments in ter-
minology and perspective.
On balance, our interactions over more than four years have
resulted in an increase in our individual collaborative strengths
and our collective realisation of the enormous potential afforded by

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:47:46, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.001
9

Is there a Future for the Past? 9

working together. We have become more adept in anticipating areas


where care must be taken and what approaches to disagreements work
best. The collaboration chapter (Chapter 8) has suggestions for ways
to resolve many common difficulties among researchers. Sometimes,
though, absolute concordance among authors is neither desirable or
possible; thus all collaborators are not involved in every chapter, and
where there are differences in opinion it is reflected in the text of
that particular chapter. Remarkably, the length of time this book has
taken is comparable with non-collaborative multi-authored books.
Several themes have been added to those that guided us earl-
ier. We have paid particular attention to issues of temporal scale; we
have more clearly embraced the complex systems approach; and we
have endeavoured to clarify the arguments for examining the past in
the service of the future and to sharpen the tools to do so. We make
no claims for having given comprehensive treatment of all the issues
and concepts in historical ecology, nor have we attempted broad geo-
graphical coverage; we simply address questions that this particular
group of scholars find central to their work and that others could
apply to their own practice and policy.
In undertaking this collaborative endeavour, we hope to stimu-
late similar efforts that would take up the lengthy list of areas we
have not addressed. A larger palette of issues and concepts can be
found in the crowd-sourced article Anthropological Contributions
to Historical Ecology: 50 Questions, Infinite Prospects (Armstrong
et al. 2017), but even this effort does not complete a description of
the research programme. Instead, we stress the incompleteness and
flexibility of the historical ecology framework which, like the places
we study, will continue to evolve.

Reference
Armstrong, C. G., Shoemaker, A. C., McKechnie, I., Ekblom, A., Szabó, P., Lane,
P. J., et al. (2017). Anthropological contributions to historical ecology: 50 ques-
tions, infinite prospects. PLoS ONE, 12(2), e0171883. doi:10.1371/journal
.pone.0171883.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:47:46, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.001
0
1

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:47:46, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.001
1

Part I Challenges: Time and


Memory

Downloaded from https://www.cambridge.org/core. University of Sussex Library, on 02 Aug 2018 at 16:51:39, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
2
1

Downloaded from https://www.cambridge.org/core. University of Sussex Library, on 02 Aug 2018 at 16:51:39, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
3
1

2 Historical Ecology and the


Longue Durée
Paul Sinclair, Jon Moen, and Carole L. Crumley

2.1 Introduction
The study of time is an integral and controversial part of most aca-
demic disciplines that lay claim to an understanding of societal
dynamics. In order to sketch the treatment of time in the emerg-
ing discipline of historical ecology, it is necessary to review the his-
tory of perceptions of time in time-sensitive disciplines (e.g. geology,
archaeology, biology, history). The aim is to examine some of the
shortcomings of the traditional approach, and finally to engage with
a relatively new vision of time that has emerged in concert with
research into complex systems.
A goal of this volume is to explore elements of a conceptual
framework for historical ecology; accordingly, we address growing
contradictions and challenges to the linear temporal framework that
undergirds Western scholarship. Often assumed is the Newtonian
world view of one linear temporal dimension and three spatial dimen-
sions. In this chapter we challenge some of the easy assumptions
about time to which many historical ecologists have long subscribed.

2.2 Studying Time in Western Disciplines


2.2.1 Geology
The development of earth history (later geology) began with
Aristotle’s (384–322 BC) observations of the slow rate of change of
the Earth relative to the human lifespan (Aristotle n.d.). For Aristotle,
time is infinite and intimately involved in change, at least in the case
of his astronomical observations of revolving spheres linked to cycli-
cal motion.

13

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
4
1

14 Sinclair et al.

From the time of Saint Augustine, in the early part of the first
millennium AD, to the Middle Ages, development of these ideas was
constrained within the limited time frame of Christian creation-
ist views. This was increasingly questioned as the study of geology
addressed the relative antiquity of physical deposits. The first pos-
tulates of stratigraphy based on observation of the accumulation of
sequential and coeval deposits were introduced in the seventeenth
century by natural scientists (e.g. Steno 1669). This milestone estab-
lished the principles of superposition and original horizontality
regarding the relative age of strata and mode of deposition.

2.2.2 Archaeology
In archaeology the study of time and its characteristics has been an
integral part of the development of the discipline since the seven-
teenth century. The famous mounds at Gamla Uppsala in Sweden
saw the first application to a site of a chain of deduction, formulated
by Olof Rudbeck the Elder (1630–1702), that allowed calculation
of the age of deposits given the depth of accumulated turf and esti-
mates of its rate of increase (King 2005). While many of his conclu-
sions were wanting, Rudbeck was a pioneer who anticipated modern
archaeological methods.
By the eighteenth century, geologists, archaeologists, natural-
ists, and philosophers were challenging the biblical interpretation of
earth history, and geology was established as an empirical science.
The concept of stratigraphy was introduced to archaeology from the
palaeontological and geological work of William ‘Strata’ Smith (1769–
1839), James Hutton (1726–97), and Charles Lyell (1797–1875). It was
Hutton who developed the idea of ‘deep time’, or geological time,
challenging the seventeenth-century ecclesiastical creationist views
that the world was only 6,000 years old. John Frere (1740–1807) and
Jacques Boucher de Perthes (1788–1868) employed stratigraphy
and geology to interpret European Palaeolithic sites that held tools
and the bones of extinct species.
It was not until the publication of Charles Lyell’s Principles
of Geology (1830) that the last remnants of catastrophism,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
5
1

Historical Ecology and the Longue Durée 15

referring to the biblical flood, were overturned and the Principle of


Uniformitarianism was established. Thereafter, processes occurring
in the past were recognised as the same as those of today, accounting
for both geological features and their formation, uniting structure
with process. This led to the formulation of a system of eons, eras,
epochs, and periods based on the increasing availability of regional
stratigraphic sequences worldwide (Berry 1987; Gibbard & van
Kolfschoten 2004).
By the late nineteenth century in Europe, a stratigraphy-based
comparative chronology was emblematic of archaeology. This broad
chronological understanding of human history was subsequently
exported, not altogether successfully, to different parts of the world,
e.g. India and Africa (Lubbock 1865; Evans 1872, 1881; Gräslund
1987; Trigger 2006; Goodrum 2008).
This essentially inductivist approach in archaeology focussed
on building relative dating sequences using comparisons of archaeo-
logical sites and artefacts. A growing chronology was later refined
with environmental proxies such as varve sequences in lakes and
tree-ring dating. It was, however, not until the major breakthrough
of radiometric dating, by Walter Libby in the 1940s, and its applica-
tion in the 1950s, that assertions of an absolute chronology became
commonplace. The conception of history archaeologists used in the
1950s and 1960s was descriptive empiricist. The flow of time was
seen as a series of events needing to be ordered relatively, in time,
using stratigraphy, comparative artefact analysis, and historical doc-
umentation, accompanied by 2D representations.
The rise of the New Archaeology in the 1970s introduced a
strong focus on societal dynamics using a positivist, deductive
approach to social and cultural hypotheses aiming to draw archae-
ology into the sciences (Clarke 1968; Watson, Le Blanc, & Redman
1971). The knowability of the past was strongly asserted and archae-
ological remains were projected in a 3D spatial universe which
continued to be calibrated with a linear concept of time. While con-
temporary archaeology has largely left the New Archaeology behind,
it has yet to revise a rather ‘time-worn’ concept of time.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
6
1

16 Sinclair et al.

2.2.3 Biology
Thinking about biology and time, and the place of humanity in
nature, also has a venerable antiquity going back to pre-Socratic
thinking. The investigative frame, derived from Aristotle with his
emphasis on motion and time, gave way to the biblically based crea-
tionist view in which all living organisms had a fixed status and tem-
poral place in the greater scheme of the natural world. Throughout
the Middle Ages, this hierarchical and essentially linear framework
did little to emphasise change, and temporal concepts remained rela-
tively undeveloped. The idea of differentiated species, put forward
by the Swedish physician and biologist Carl Linnaeus (1735), were
formulated within a creationist view of the world.
A less ecclesiastical and more scientific view of the natural
world emerged throughout the eighteenth and early nineteenth cen-
turies (e.g. Maupertuis, Buffon, Lamark, Cuvier). It was not until
the mid-nineteenth century, however, through the publication of
Darwin’s On the Origin of Species (1859), that evolution became
widely accepted across geology, archaeology, and biology. The use of
a vast array of proxy-data sources in the natural sciences is now fun-
damental for studies of human-environment interactions.

2.2.4 Ecology
The early- to mid-twentieth-century development of ecology, con-
cepts of the biosphere, and human ecology paved the way for the
development of systems ecology, an important conceptual focus for
the practice of historical ecology (Vernadsky 1929; Hutchinson 1957–
93; Fraser Darling 1969; see also Odum & Odum 1953). Once again,
natural science data were projected in 2D, then later in 3D, and a
linear concept of time was used to calibrate the variables. The time
dimension was, however, addressed by the ecologist Holling (1973),
with his focus on the natural cycle and resilience theory, and was fur-
ther elaborated into the concepts of panarchy (Gunderson & Holling
2001) and spatial resilience (Cumming & Norberg 2008).

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
7
1

Historical Ecology and the Longue Durée 17

2.3 The Search for New Concepts of Time


The bulk of contemporary Western research in ecology, archaeology,
history, human geography, and much of biology is anchored in a linear
perspective, with exceptions (e.g. Hägerstrand 1970, 1983; Carlstein
1982; DeLanda 1997). Cyclical and multi-scalar perspectives have
been considered, especially in relation to Chinese and Indian sci-
entific traditions (Needham 1954–9; Murray 1999). The concept of
time has been expanded by developments in ecology (Peterson 2002;
Bengtsson et al. 2003; Brand 2005; Sun & Ren 2011), in several social-
science disciplines through memory and ethnographic studies (e.g.
Barthel, Crumley, & Svedin 2013a, 2013b), and in creative writing
(e.g. science fiction, Robinson 2009).
The first attempts to move away from conceptualising his-
tory as a series of events towards a distinction between different
time scales came in the early decades of the twentieth century. The
German historian and philosopher Oswald Spengler (1880–1936)
wrote in his influential The Decline of the West (1918):

I see in place of that empty figment of one linear history . . . the


drama of a number of mighty cultures . . . each having its own
life . . . its own death . . . Each culture has its own new
possibilities of self-expression, which arise, ripen, decay and
never return . . . I see world history as a picture of endless
formations and transformations, of the marvellous waxing and
waning of organic forms. The professional historian, on the
contrary, sees it as a sort of tapeworm industriously adding onto
itself one epoch after another.
(emphasis by eds.)

In 1920s France, a group of young historians broke with their dis-


cipline by emphasising the context of events, i.e. their geography,
ethnography, folklore, geology, and climate. Two of them, Marc
Bloch and Lucien Febvre, began an influential journal initially called
Annales d’histoire économique et sociale (today Annales: Histoire,
Sciences Sociales).

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
8
1

18 Sinclair et al.

The Annalistes’ humanistic and dialectical approach provided


a new collaborative framework, especially for the study of local and
regional history. Their analysis of processes that vary along interdigi-
tating temporal scales of duration, intensity, and periodicity proved
especially fruitful. They conceived of historical processes in terms
of événement (event), conjoncture (their cultural and historical
context), and longue durée (long-term trends). Interpretation relies
on all three time frames: an account must be set in immediate, as
well as more distant contexts, whether found in a cleric’s records of
famine victims or deep in a glacier half a world away.
In his book La Méditerranée et le Monde Méditerranéen a
l’époque de Philippe II (1949) (published in English in 1972 as The
Mediterranean and the Mediterranean World in the Age of Philip II),
Fernand Braudel (1902–85) portrayed the Annaliste conception of time
in history as a multiplicity of rhythms and tempi. He argued that the
most immediately perceptible – that which transfixed traditional nar-
rative historians – was the short time span, the time of days, months,
and years (événement). Here, the focus was on the actions of individu-
als and the history of events: ‘a surface disturbance, the waves stirred
up by the powerful movement of tides’. A second type of historical
time span (conjuncture) was one whose duration could be measured
in decades and centuries, and which was concerned with the cyclical
rise and fall of ‘economies and states, societies and civilizations’. But
underlying all these was yet another time span – what he termed the
longue durée – ‘which exists almost out of time and tells the story of
man’s contact with the inanimate’ (see Gould 2002).
Historical time for Braudel, then, was distinguished by the sim-
ultaneous presence and interaction of these three levels: geographical
time, social time, and individual time. But in the histories he wrote,
Braudel made clear that it was the longue durée that constituted the
most meaningful context through which to examine and understand
history (Tomich 1958; Braudel 1972; Kidambi 2011).
Much has been written about the Annales School and espe-
cially Braudel, evoking, in particular, the privileging of the role of

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
9
1

Historical Ecology and the Longue Durée 19

environmental variables, the incipient determinism at the heart of


the longue durée, and the lack of a dynamic intersection of temporal-
ities (e.g. McGlade 1999: 146–7). Others have focussed on the indubi-
table utility of the integrative landscape, region, culture, and climate
approach that is the hallmark of this School (Bintliff 1991; Knapp
1992). What is clear, however, is that the Annalistes have carried the
discussion of time in history and archaeology forward into current
conceptions of non-linear time.
It is difficult to maintain that, as a community of practice, the
Annalistes have managed to overemphasise environmental condi-
tions, given the great paucity until recently of non-Annaliste his-
torical treatments that include any environmental parameters. In
fact, it is the Annaliste Emmanuel Le Roy Ladurie who introduced
several generations of historians to the richness of an analysis that
blends weather, climate, and history (1988; published in 1967 as
Histoire du climat depuis l’an mil). Regarding the charge of longue
durée determinism, one may level this at Braudel, but not at the
majority of Annalistes. This leaves two significant issues which
have not yet been widely addressed: McGlade’s critique of the miss-
ing ‘dynamic intersection of temporalities’, and the looming cri-
sis in causation (1995, 1999), which will be addressed later in this
chapter.
A significant advantage of the later twentieth-century post-
structuralist and post-colonial thinking about the world’s societies
is an increased appreciation for what can be called ‘emic time’. This
refers to the different approaches for conceptualising the duration
and passage of time, whether cyclic or linear, in hunting, pastoral,
agricultural, and urban societies throughout the world. Inspired by
the famous study of the Nuer (Evans-Prichard 1954), many other
concepts of time are now well known. Further development of post-
structural analyses, inspired by Anthony Giddens (1979), puts a
spotlight on the role of the cognisant individual in shaping human
history, and a much more sophisticated treatment of agency and cau-
sality in human socio-cultural interactions.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
0
2

20 Sinclair et al.

The variety of temporal cosmologies should not be underesti-


mated, and it is a challenge to integrate alternative epistemological
points of departure. The divergences from modernist views of time’s
arrow within privileged forms of emic time can be illustrated with
some central African pre-capitalist perceptions of time as ‘being in
the present and becoming the past’ (Mbiti 1990). From this perspec-
tive, the future tense can be seen as a modernist imposition. Lucas
(2005) differentiates between past actors’ perceptions of time and
those of archaeologists; in a sense, this can be characterised as emic
and etic time.
Increasingly, expanded temporal frameworks guide many his-
torical sciences and conform to a holistic view of the multivalent
Earth system. Geological interpretation, for example, deftly com-
bines structures (e.g. crystallography, stratigraphy) and processes
(weathering, uplift), linking scales of time and space. In A Thousand
Years of Non-linear History DeLanda (1997) attempts to escape the
limitations of fluid time. He points to uneven, rapid change and peri-
ods of stagnation (punctuated time, accelerating and decelerating
rates of change) across many fields of study, from geology, biology,
and climate to the study of urban trajectories. Gould and colleagues’
(Gould 2002) view is of punctuated equilibria, where the evolution of
species, seen in geological time, is characterised by periods of rapid
evolution followed by long periods of stasis. To reconcile this with
the traditional view of evolution as a series of small steps, Gould
introduces three different tiers of time. The first consists of a slow,
gradual change as a response to natural selection pressures in ecologi-
cal time. The second is geological time characterised by sequences of
rapid and slow change, while the third comprises ‘deep time’ where
mass-extinction events shape evolutionary patterns. This conception
of dynamic material reality points to the emerging consensus around
the precepts of complex systems science.
The disciplines of history and archaeology have yet to deal
with the fundamental shift from the Newtonian model, in which
objects are located in three-dimensional space and move through

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
1
2

Historical Ecology and the Longue Durée 21

time according to the laws of thermodynamics. This view holds only


the present to be real, and so is dubbed ‘presentist’ by philosophers.
An alternatively named ‘growing block time’ holds both past and pre-
sent to be real, while the future is indeterminate. These concepts of
time are both consonant with our ordinary perception of the passage
of time, measurement of change, and narrative of historical events
(Davies 2014). In the wake of Einstein’s theories of relativity and the
development of quantum physics, however, our ordinary perception
is overturned. Space and time are enfolded into a continuum, so that
past, present, and future are part of the space-time manifold, the
‘eternalist’ view of time. The measurement of change is dependent
on the observer’s frame of reference. Events and their measurement
are entangled. Linear causality is but one mode of causality in a com-
plex universe.
At the beginning of the twentieth century the philosopher
Bertrand Russell and the mathematician A. N. Whitehead devel-
oped an approach to complexity as nested logical types, each gov-
erned by its own logic, none reducible to another. In Steps to an
Ecology of Mind (1972) and Mind and Nature (1979), the anthropol-
ogist Gregory Bateson applied this approach to human-environment
interrelations. The extension of this application to the current
understanding of non-linear systems – replacing nested spatial
hierarchies and linear temporalities – is the task facing the histori-
cal disciplines if they are to deal with the complexities of a non-
Newtonian world.

2.4 Investigating Time in Complex Systems


Complexity science (CS) is the study of dynamic non-linear systems,
i.e. systems that are not in equilibrium and do not act in a predict-
able manner. This includes the poly-disciplinary study of complex
adaptive systems (CAS), which are dynamic and non-linear (not
temporally sequential) systems. The origins of CS are in physics and
biology, through the study of chaos and non-linear dynamics (Gleick
1987; Parrot & Lange 2013).

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
2

22 Sinclair et al.

In CAS, complex relates to a dynamic network of interactions


among elements that are not static, and where the behaviour of the
ensemble cannot be predicted by the behaviour of the components.
A CAS has no overarching hierarchy of elements, but rather a complex
heterarchy of interacting elements that may sometimes dominate
the system, and at other times be subordinate to it (Crumley 2005).
CAS are adaptive in that elements can self-organise in response to
changing conditions. They are systems in that elements form inter-
connected networks.
While the concept of a system may resonate well with natural
scientists, some researchers within social sciences and humanities
are less attracted, claiming that systems ontologies or boundaries are
seldom definable (Olsson et al. 2015). While this is essentially true,
CAS systems are simply defined as ‘a set of things – people, cells,
molecules, or whatever – interconnected in such a way that they pro-
duce their own pattern of behavior over time’ (Meadows 2008). This
definition allows the description of many categories (e.g. species and
ecosystems, the spread of diseases, traffic networks, neighbourhood
organisations, and financial markets).
Several features distinguish contemporary complex-systems
thinking from earlier systems theory, which, following Clements
(1928), assumed that ‘natural’ systems could be modelled with a few
key variables and would return to equilibrium after being disturbed.
Reflecting rejection by ecologists of the Clementsian idea, Levin
(1998) defines ecosystems as complex adaptive systems in which the
interaction of life processes forms self-organising patterns across dif-
ferent temporal and spatial scales.
While both mid-twentieth-century systems theory (roughly
1930s to 1970s) and the new complex-systems thinking address
the organisation of information, there are important contrasts.
The earlier paradigm – a cornerstone of the New Archaeology dur-
ing the 1960s and 1970s – held the tantalising possibility for many
archaeologists that a predictive science of human behaviour could
be constructed in the language of mathematics and philosophy

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
3
2

Historical Ecology and the Longue Durée 23

(Binford & Binford 1968; Watson et al. 1971; Flannery 1972). Parallel
trends developed in ecology and elsewhere in the biological sciences
(Ellen 1982). Contemporary CAS, however, promises no familiar path
for causation, much less for predictive modelling.
Moreover, CAS research is not itself a single theoretical orienta-
tion, but a transdisciplinary aggregate of several strands of investiga-
tion, widely applied in the biological, physical, and social sciences.
It is now recognised that an amalgam of methods and perspectives is
necessary for a holistic understanding of interactions among many
elements/actors. In an example from the social sciences, following
Agar (1973, 2004, 2007), Bourgois and colleagues (1997), and others,
Meadows (2008) claims that hunger, poverty, environmental degrada-
tion, drug addiction, and war can be seen as problems that require an
understanding of the often surprising links between them. No one
deliberately creates a problem, no one wants it to persist, but it results
in undesirable behaviours. To address such issues, the interaction of
diverse elements – each with its own history – must be understood.
CAS brings several relevant concepts to the study of time, such
as non-linearity, initial conditions, emergence, basins of attraction,
and path dependence. Particularly important for historical ecology is
a focus on scale and on time as it relates to causation. These intrigu-
ing ideas can broaden the study of combined social and environmen-
tal change across time and space, and take it into the future (van der
Leeuw & McGlade 1997; van der Leeuw 1998; Beekman & Baden
2005; Crumley 2005; Redman 2005; Lane et al. 2009).
As ecologist Evelyn Pielou (1975) demonstrated several decades
ago, scale defines diversity. Scale relates both to how systems organise
themselves and to how they are studied. Scale in all its dimensions –
spatial, temporal, economic, social, political, spiritual – is of central
importance. If the wrong analytic scale is chosen, critical evidence of
similarity or difference may go undetected. Systems organise them-
selves and evolve at multiple scales. To follow changes over time, dif-
ferent types of shifting scales, both analytic (epistemic) and systemic
(ontologic, inherent), must be employed.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
4
2

24 Sinclair et al.

In order to understand change in the past, the traditional


approach is to establish a temporally ordered chain of events,
termed causation, which allows generalisation about a behaviour.
Causation is a relationship between events, properties, variables, or
states of affairs and implies a relationship of dependency between
the cause and the effect. Since complex systems (such as the human-
environment relation) are comprised of both linear (predictable) and
non-linear (emergent) properties, the establishment of causation is
both fundamental – if we are to learn from the past – and problematic.
Since significant events (thresholds) in complex systems are
the sum of many other events/properties that operate at various tem-
poral and spatial scales and different dimensions (system history),
the ability to establish a clear line of causation (A causes B) is com-
promised. It is, however, possible to analyse change in a system by
identifying variables that concatenate to bring change about.
This requires a meta-theoretical approach which takes into
account the properties of dynamical systems that include human
societies. Counterfactual, probabilistic, derivation, manipulation,
and process theories are only some of the means by which scholars
have approached this problem. Historical ecology, deeply rooted in
the soil of places and their history, offers a means to address multiple
causation in the formation and evolution of landscapes.
Traditionally, research has addressed complicated problems by
breaking the whole into smaller parts that are more easily studied.
Like chronological time, this reductionist process has served scholar-
ship well over the past centuries, and has greatly increased human
scientific knowledge. Reductionist science avoids higher-level phe-
nomena and patterns by assuming that they can be described by pro-
cesses at lower levels. In contrast, complexity science (CS) focuses
directly on the higher-level phenomena, or so-called emergent pat-
terns. CS is thus less concerned with objects, agents, or states and
more with how these entities interact, the changes that follow, and
the recognition of overall patterns at many scales. The result is a
holistic, encompassing approach to complexity.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
5
2

Historical Ecology and the Longue Durée 25

Systems and interactions are ubiquitous. ‘Clearly, interactions


are central to making a system complex. Without interactions, there
is no emergence, no feedback and no cross-scale linkages’ (Parrot &
Lange 2013). By focussing on these interactions, complexity science
may represent a new way of looking at old problems, and a new way
of analysing old data. Accordingly, we argue that complex systems
analysis is an important component of historical ecology.

2.5 Properties of Complex Systems


How can we recognise a complex system? While the definitions
given earlier in this chapter may seem a bit vague, there are several
characteristics of complex systems that may guide us. This section
is based on Johnson (2007) and Parrott and Lange (2013), unless oth-
erwise stated.

• Connectivity: There are many interacting agents, such as people, animals,


plants, or molecules. The interactions can result from close proximity,
where the agents form groups, or be based on the exchange of information
or other commodities. Often these interactions create different kinds
of network, and network analysis has become an important part of
complexity science (Webb & Bodin 2008).
• History: Interactions are influenced by path dependency, memory, or
feedback. Path dependency means that past events amplify through
positive feedback to strongly affect interactions today. For instance,
the loss of Finland to Russia in the Swedish-Russian war of 1808–9 led
to border closures, which influenced, and continue to influence, the
development of reindeer husbandry in Sweden (Moen & Keskitalo 2010).
Memory may also affect interactions (see Chapter 5 for examples of how
memory may affect conservation actions). Feedbacks are chains of events
that influence themselves, either positively or negatively. A classic
example of a positive feedback is when higher temperatures caused
by climate change melt ice and snow at high latitudes, changing the
albedo (the reflectance of the surface), trapping more heat, which in turn
increases temperatures, and melts more snow, etc. All these interaction
modifiers will have strong effects on the dynamics of a system (see later
in this chapter). The existence of feedback loops also means that agents

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
6
2

26 Sinclair et al.

can adapt their strategies and behaviours, based on history or earlier


experiences, and that systems can change over time.
• Context: A system is typically open, meaning that it can be influenced by
its larger context. It also means that any system boundary is more or less
arbitrary. This may seem trivial since the only truly closed systems exist
in the laboratory, but it is important because most of the theory describing
system dynamics in physics and mathematics is based on closed systems
(since they are mathematically easier to analyse). Real-world systems,
however, have more complex behaviours than theoretical systems.
• Emergence: Systems often show surprising emergent phenomena. Because
of feedback loops and the openness of the systems, they are seldom or
never in equilibrium, i.e. they are not static or stable. As agents respond
to one another or to changes in their environment, almost any pattern
can appear. One example of a surprising emergent phenomenon is a black
swan stock market crash, seemingly impossible to predict based on the
behaviour of individual traders (especially since some of these may be
computers that act on information in microseconds). Such emergent
phenomena seem to arise spontaneously since they occur without
any central controller, but as a consequence of all the interactions and
adjustments of the system. Other examples are the beautiful flowing
patterns of bird flocks or fish schools, caused by the simple rule of ‘avoid
bumping into your neighbour’ (Figure 2.1).

2.6 Time in Complex Systems


Time is inherent in the dynamics of complex systems. Its presence
is often termed thresholds, tipping points, regime shifts, or critical
transitions, which all describe non-linear responses to changes in
the environment. Such abrupt changes in system dynamics create a
sort of punctuated rhythm to time. The study of non-linear dynamics
draws heavily on precursors in physics, mathematics, and ecology, for
instance on the concept of alternative stable states (e.g. May 1977),
but is now often applied to environmental and societal contexts. The
basis for such discussions is how the focal system responds to chan-
ging conditions.
A tipping point can be defined as the critical juncture in a
situation, process, or system beyond which a significant and often

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
7
2

Historical Ecology and the Longue Durée 27

Figure 2.1 A wedge of starlings.


Photo: Paolo on Wikimedia Commons. CC-BY-SA.

unstoppable effect or change takes place (Merriam-Webster 2017).


While Malcolm Gladwell (2000) makes good contemporary use of
the term, its origins are not value neutral. In the 1950s, the phrase
‘tipping point’ was used to establish the percentage of non-whites in
a neighbourhood that would generate white flight (Merriam-Webster
2017). This should act as a reminder that change, however it is
brought about, and especially if it is rapid, is never neutral.
Employing the more current use of the term, Gladwell argues
that tipping points in social contexts are governed by three different
rules, or agents of change: the Law of the Few, the Stickiness Factor,
and the Power of Context. The Law of the Few refers to people with
very special social skills. They could, for instance, be persuasive or
very connected entrepreneurs, which invokes network theory (see
e.g. Webb & Bodin 2008). The Stickiness Factor refers to the way
information is packaged to attract people’s attention. The Power of
Context means that the community or society has to be ready to
receive the information or idea, and that the same idea may or may

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
8
2

28 Sinclair et al.

not spread, depending on the context. Many of Gladwell’s ideas can


be found in the study of epidemics (e.g. Merrill 2012) and the spread
of innovations (Barnett 1953; Rogers 2003). In the latter case, the
focus is on who the early adopters are, and how a critical number has
to be reached before an innovation begins to spread, usually described
as around 16 per cent of the population.
Many enthusiasts searching for tipping points cling to older
ideas about causation, since it is thought that systems could be man-
aged to reduce the risk of passing undesirable thresholds or tipping
points (Werner et al. 2013). Considerable research has been devoted
to finding ‘early-warning indicators’ (e.g. Pace et al. 2015). The detec-
tion of different kinds of early-warning signals may be possible, but
this depends on the type of non-linear dynamics involved (Lenton
2013; Pace et al. 2015). Some research has focussed on statistical
behaviour of time-series data, including what is known as critical
slowing down, i.e. the system becomes more ‘sluggish’, exhibits
increased variability or changes in data symmetry close to the tipping
point (Carpenter et al. 2011). All these methods require rather a lot of
data, and seem to have been applied only ex post facto and have con-
sequently not predicted any abrupt changes. The search for indicators
is difficult as the temporal and spatial scales involved often make it
impossible to experimentally manipulate the systems.
Searching for early-warning indicators by selecting histori-
cally documented systems that have experienced abrupt changes
may, however, also introduce a statistical error (Boettiger &
Hastings 2012). Nevertheless, the continued understanding of non-
linear dynamics, and the search for ways of predicting whether
systems are at risk of abrupt change is an important, but very chal-
lenging part of complex systems study. History could provide clues,
but first it may be necessary to recognise the always partial analytic
universe of relevant system elements, as well as the shortcomings
of a purely statistical approach. There has been some discussion of
meta-data analysis of a ‘perfect storm’ variety, where many histori-
cal examples exhibit similar tendencies (e.g. polities experiencing

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
9
2

Historical Ecology and the Longue Durée 29

instability due to cascading or simultaneous drought, economic


shifts, and conflict).
It is perhaps worth considering that, in systems with a human
component, tipping points occur too late in time; by then our atten-
tion is drawn to the rapid systemic change, rather than an inspec-
tion of earlier decision nodes that may have precipitated the change.
A focus on decision nodes recognises the multiplicity of system
futures, which are diminished once the decision is put into practice.
This is not to say that change in human systems is always due to bad
decisions, because it is quite possible that whatever the decision, the
system might have ‘flipped’ anyway. The point is that, in order to
learn from the analysis, the context of the decision (e.g. what infor-
mation was not available) would offer greater insight.
Soranno and colleagues (2014) suggest that a first step in study-
ing complex systems could be to construct a conceptual model based
on the current understanding of the system and the drivers affecting
it. This can be followed by building a database with the available
information and finally the translation of this conceptual model into
a set of models that represent competing hypotheses of how the sys-
tem works. While this may be a powerful approach in some cases,
there will be many situations for which data are missing or exist only
in some qualitative form. In these cases other approaches, such as
various scenario techniques, fuzzy cognitive maps (Grey et al. 2015),
or Stommel diagrams (Scholes et al. 2013) are required. A good exam-
ple is Henry Stommel’s 1963 diagram, now an important tool in ecol-
ogy, of a three-dimensional graphical depiction of four-dimensional
phenomena (Haury et al. 1978; Levin 1992; Schneider 2001; Vance &
Doel 2010).
If we accept that the future is not determined by some higher
power, the number of futures is infinite at each particular time. The
future that actually emerges is the result of many decisions, both
conscious and unconscious, and other events, some of which are
controllable, while others are not. Homer-Dixon and colleagues
(2015) argue that critical changes to a system (what they refer to as

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
0
3

30 Sinclair et al.

‘synchronous failures’) are characterised by simultaneous multiple


stresses on a system. If the system is connected enough, this may
result in ramifying cascades of effects that change the system from
one state to another. They also stress that to study these critical
changes, researchers should focus on the complex, dense, and evolv-
ing causal links in humanity’s energy, e.g. food, water, climate, and
financial systems. Holistic approaches of this type are familiar in his-
torical ecology.
Homer-Dixon and colleagues also suggest that current crises
are very different from those of the past, as today’s systems are more
global and closely connected and therefore have no unaffected parts
that can adapt or transform to respond to the crisis (whether environ-
mental or financial). The researchers argue that three aspects make
current crises different: a dramatic increase in the scale of economic
activity, increasing homogeneity of practices, institutions, and tech-
nologies, and finally a rapid increase in connectivity. We argue that
this is a quantitative difference rather than a qualitative one. While
past crises may not have had global consequences, many scales of
economic and institutional activity, dynamic environments, and pat-
terns of connectivity have always been important. Examples include
the collapses of both the Roman and the Mayan polities (Tainter &
Crumley 2007; Chase & Scarborough 2014).

2.7 Advantages of Complex Systems in


Historical Ecology
The combination of multiple stresses, interactions of many agents,
connectivity, networks, and multiple scales of influence means that
there will never be one single version of the past that explains the
road we have taken to get where we are today. The relative impor-
tance of events will change depending on the perspective. It may even
be that events and stresses are so interdependent that they cannot be
separated or given different weights in explanations.
A good metaphor is found in physics, where the physical phe-
nomenon of quantum entanglement affects the interaction of pairs,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
1
3

Historical Ecology and the Longue Durée 31

or groups of particles, so that the quantum state cannot be ascribed


to the particles independently, but only to the system as a whole.
While this undoubtedly complicates any historical study, it also frees
up the mind to consider multiple lines of reasoning and how these
can inform the future. After all, history gives us the only completed
experiments in complex systems, and the past should be explored to
the fullest in order to inform current challenges.
Complexity science gives a particular focus to different kinds
of critical transitions. We suggest that valuable lessons for the future
would be to focus on historical periods of instability rather than
periods of relative stability (for some popular, not always successful
attempts see Turchin 2003; Diamond 2005). While stability may be
a measure of sustainability over time, studies of the conditions sur-
rounding transitions, and the characteristics of such shifts, may be
just as important, both to avoid unwanted transitions and to achieve
desired transformations (Tainter 2006; Scheffer 2009).

2.8 New Frameworks for Time


We are not arguing that historical ecologists should abandon linear
time with its intendant logic of cause-and-effect. Instead, we advo-
cate an increased awareness of the kinds of time that have made
our world and the multiplicity of potential causal constituents of
observed or experienced effects. If linear time, reckoning from geo-
logical strata and the earliest calendrical systems to this morning’s
wake-up alarm, is epistemological (schemes invented to measure),
then the reckoning of time in complex systems is ontological, a con-
dition beyond human construction and a characteristic of the uni-
verse itself (Kauffman 1993, 1995).
If this is so, how are we to integrate the reality of varves, spe-
leothems, and tree rings with an understanding that time is relative?
Perhaps geology can lead the way by asking researchers and planners
to think, simultaneously, about structure and process. At the most
basic level, geology is both a spatial and a temporal science. In order
to understand how the planet (and the universe) have changed, tools

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
2
3

32 Sinclair et al.

that temporally order events must be used. Geological inquiry, how-


ever, does not stop there; it is equally important to understand the
drivers of change in the system as a whole, which in turn requires
complex-systems thinking.
As we must use multiple lines of reasoning when studying
complex systems or practising historical ecology, it follows that we
also embrace multiple views of time. Time is perhaps not always
relative, and each process moves at a different and variable rate. As
Bailey notes: ‘differing timescales bring into focus different features
of behaviour, requiring different sorts of explanatory principles’
(1981: 103; 2007: 200).
McGlade (1999: 156) asserts that ‘it is not change per se that
is important, rather . . . we must shift our focus to questions which
deal with (i) the rate of change and, perhaps more important, (ii)
the change in the rate of change. It is these attributes which above
all define system dynamics’ (see also Garnsey & McGlade 2006).
Multiple timelines with different rates of change give a view of time
as a ‘braid’ rather than as a line. Processes move at different rates,
and different perspectives and scales give distinct views of time (e.g.
Gould’s three tiers). A useful metaphor is to see time as a braided
river, where channels flow at different speeds (see Figure 2.2).
Another insight from McGlade is that ‘changing rates of
change . . . constitute the structural history of societies, and which
in the final analysis produce a model of history as a non-linear
dynamical system’ (1999: 147). As organisations such as the Islamic
State of Iraq and the Levant (ISIS/ISIL) demonstrate, ‘power flows
in many channels’ (Samford 2007) and can manifest itself entirely
outside the framework of nation-states and beyond their control
(e.g. Anderson 1983; Crumley 2005; Scott 2009). It does not mat-
ter whether such work is termed ‘the generation of competing
hypotheses about system function’, ‘the analysis of the dynamic
interplay of humans and environments’, or ‘the construction of sce-
narios’. Historical ecology can offer new ways to understand our
complex world.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
3

Historical Ecology and the Longue Durée 33

Figure 2.2 Time as a braided river with different processes flowing at


different rates.
Photo: Avenue on Wikimedia Commons. CC-BY-SA.

References
Agar, M. (1973). Ripping and Running: A Formal Ethnography of Urban Heroin
Addicts. New York: Seminar.
Agar, M. (2004). We have met the other and we’re all nonlinear: ethnography as a
nonlinear dynamic system. Complexity, 10(2), 16–24.
Agar, M. (2007). Dope Double Agent: The Naked Emperor on Drugs. Lulu.com.
Anderson, B. (1983). Imagined Communities: Reflections on the Origin and
Spread of Nationalism. New York: Verso.
Aristotle. Meteorology. Book 1, Part 14.
Bailey, G. (1981). Concepts, time-scales and explanations in economic pre-
history, in Economic Archaeology, A. Sheridan and G. Bailey, eds., British
Archaeological Reports, pp. 97–117 (International Series) 96, Oxford.
Bailey, G. (2007). Time perspectives, palimpsests and the archaeology of time.
Journal of Anthropological Archaeology, 26, 198–223.
Barnett, H. G. (1953). Innovation: The Basis of Cultural Change. New York:
McGraw-Hill.
Barthel, S., Crumley, C., & Svedin, U. (2013a). Bio-cultural refugia: safeguarding
diversity of practices for food security and biodiversity. Global Environmental
Change, 23(5), 1142–52.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
4
3

34 Sinclair et al.

Barthel, S., Crumley, C., & Svedin, U. (2013b). Bio-cultural refugia: combating the
erosion of diversity in landscapes of food production. Ecology and Society,
18(4), 71.
Bateson, G. (1972). Steps to an Ecology of Mind. New York: Random House.
Bateson, G. (1979). Mind and Nature – A Necessary Unity. New York: Bantam Books.
Beekman, C. S. & Baden, W. S., eds. (2005). Nonlinear Models for Archaeology and
Anthropology. Aldershot (Hampshire), UK: Ashgate Press.
Bengtsson, J., Angelstam, P., Elmqvist, Y., Emanuelsson, U., Folke, C., Ihse, M.,
Moberg, F., & Nyström, M. (2003). Reserves, resilience and dynamic land-
scapes. Ambio, 32(6), 389–96.
Berry, W. B. N. (1987). Growth of a Prehistoric Time Scale: Based on Organic
Evolution. Palo Alto, CA: Blackwell Scientific Publications.
Binford, S. & Binford, L., eds. (1968). New Perspectives in Archaeology.
Chicago: Aldine.
Bintliff, J., ed. (1991). The ‘Annales’ School and Archaeology. New York: New York
University Press.
Boettiger, C. & Hastings, A. (2012). Early warning signals and the prosecutor’s fal-
lacy. Proceedings of the Royal Society B, 279, 4734–9.
Bourgois, P., Lettiere, M., & Quesada, J. (1997). Social misery and the sanctions of
substance abuse: confronting HIV risk among homeless heroin addicts in San
Francisco. Social Problems, 44, 155–73.
Brand, F. (2005). Ecological Resilience and its Relevance within a Theory of
Sustainable Development. UFZ Centre for Environmental Research Leipzig-
Halle in the Helmholtz-Association. Department of Ecological Modelling
UFZ-Report 0 3/2005 ISSN 0948-9452.
Braudel, F. (1972). The Mediterranean and the Mediterranean World in the Age of
Philip II. Originally published in 1949. New York: Harper.
Carlstein, T. (1982). Time Resources, Society, and Ecology: On the Capacity for
Human Interaction in Space and Time. London; Boston: Allen & Unwin.
ISBN 0043000827. OCLC 7946554.
Carpenter, S. R., Cole, J. J., Pace, M. L., Batt, R., Brock, W. A., Cline, T., Coloso,
J., Hodgson, J. R., Kitchell, J. F., Seekell, D. A., Smith, L., & Weidell, B. (2011).
Early warnings of regime shifts: a whole-ecosystem experiment. Science, 332,
1079–82.
Chase A. F. & Scarborough, V. L., eds. (2014). The Resilience and Vulnerability
of Ancient Landscapes: Transforming Maya Archaeology through IHOPE.
Archaeological Papers of the American Anthropological Association, no. 24.
Hoboken, NJ: Wiley.
Clarke, D. L. (1968). Analytical Archaeology. London: Methuen.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
5
3

Historical Ecology and the Longue Durée 35

Clements, F. E. (1928). Plant Succession and Indicators. New York: H. W. Wilson.


Crumley, C. L. (2005). Remember how to organize: heterarchy across disciplines.
In C. S. Beekman and W. S. Baden, eds., Nonlinear Models for Archaeology
and Anthropology, pp. 35–50. Aldershot (Hampshire), UK: Ashgate Press.
Cumming, G. S. & Norberg, J. (2008). Scale and complex systems. In J. Norberg
and G. S. Cumming, eds., Complexity Theory for a Sustainable Future, pp.
246–76. New York: Columbia University Press.
Darwin, C. (1859). The Origin of Species by Means of Natural Selection.
London: John Murray.
Davies, P. (2014). That mysterious flow. Scientific American: A Matter
of Time, 23, 8–13. Published online: 23 October 2014. doi:10.1038/
scientificamericantime1114-8.
DeLanda, M. (1997). A Thousand Years of Nonlinear History. New York: Zone
Books, New York.
Diamond, J. (2005). Collapse – How Societies Choose to Fail or Succeed.
New York: Viking Penguin.
Ellen, R. (1982). Environment, Subsistence and System: The Ecology of Small-
Scale Social Formations. Cambridge: Cambridge University Press.
Evans, J. (1872). The Ancient Stone Implements, Weapons and Ornaments of
Great Britain. New York: D. Appleton and Company.
Evans, J. (1881). The Ancient Bronze Implements, Weapons, and Ornaments of
Great Britain and Ireland. London: Longmans Green & Co.
Evans-Pritchard, E. E. (1954). Nuer Religion. Oxford: Oxford University Press.
Flannery K. V. (1972). The cultural evolution of civilizations. Annual Review of
Ecology and Systematics, 3, 399–426.
Fraser Darling, F. (1969). The Impact of Man on the Biosphere. UK: BBC Reith
Lectures.
Garnsey, E. & McGlade, J., eds. (2006). Complexity and Co-evolution, Continuity
and Change in Socio-economic Systems. Cheltenham: Edward Elgar.
Gibbard, P. & van Kolfschoten, T. (2004). The Pleistocene and Holocene epochs. In
F. M. Gradstein, J. G. Ogg, & A. G. Smith, eds., A Geologic Time Scale 2004.
Cambridge: Cambridge University Press.
Giddens, A. (1979). Central Problems in Social Theory: Action, Structure and
Contradiction in Social Analysis. London: MacMillan Press.
Gladwell, M. (2000). The Tipping Point. Boston: Little, Brown and Company.
Gleick, J. (1987). Chaos – Making a New Science. London: Sphere Books.
Goodrum, M. R. (2008). Questioning thunderstones and arrowheads: the problem
of recognizing and interpreting stone artifacts in the seventeenth century.
Early Science and Medicine, 13(5), 482–508. doi:10.1163/157338208X345759.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
6
3

36 Sinclair et al.

Gould, S. J. (2002). The Structure of Evolutionary Theory. Cambridge, MA: Harvard


University Press.
Gräslund, B. (1987). The Birth of Prehistoric Chronology: Dating Methods
and Dating Systems in Nineteenth-Century Scandinavian Archaeology.
Cambridge: Cambridge University Press.
Grey, S. A., Gray, S., De Kok, J. L., Helfgott, A. E. R., O’Dwyer, B., Jordan R., &
Nyaki, A. (2015). Using fuzzy cognitive mapping as a participatory approach to
analyze change, preferred states, and perceived resilience of social-ecological
systems. Ecology and Society, 20(2), 11.
Gunderson, L. H. & Holling, C. S. (eds.). (2001). Panarchy: Understanding
Transformations in Human and Natural Systems. Washington, DC: Island
Press.
Hägerstrand, T. (1970). What about people in regional science? Papers of the
Regional Science Association, 24(1), 6–21. doi:10.1007/BF01936872.
Hägerstrand, T. (1983). In search for the sources of concepts. In A. Buttimer, ed.,
The Practice of Geography, pp. 238–56. London, New York: Longman.
Haury, L. R., McGowan, J. A., & Wiebe, P. H. (1978). Patterns and processes in
the time-space scales of plankton distribution. In J. H. Steele, ed., Spatial
Pattern in Plankton Communities, pp. 277–327. New York: Springer
Science+Business Media.
Holling, C. S. (1973). Resilience and stability of ecological systems. Annual
Review of Ecology and Systematics, 4, 1–23.
Homer-Dixon, T., Walker, B., Biggs, R., Crépin, A.-S., Folke, C., Lambin, E. F.,
Peterson, G. D., Rockström, J., Scheffer, M., Steffen, W., & Troell, M. (2015).
Synchronous failure: the emerging causal architecture of global crisis. Ecology
and Society, 20(3), 6.
Hutchinson, G. E. (1957). Concluding remarks. Cold Spring Harbor Symposia on
Quantitative Biology, 22, 415–27. Reprinted (1991): Classics in Theoretical
Biology. Bulletin of Mathematical Biology, 53, 193–213.
Hutchinson, G. E. (1957–93). A Treatise on Limnology. Vol. I Geography, Physics
and Chemistry (1957) Wiley. Vol. II Introduction to Lake Biology and the
Limnoplankton (1967) Wiley. Vol. III Limnological Botany (1975) Wiley. Vol.
IV The Zoobenthos (1993). New York: Wiley.
Johnson, N. (2007). Simply Complexity – A Clear Guide to Complexity Theory.
Oxford: Oneworld, Oxford University Press.
Kauffman, S. (1993). The Origins of Order: Self Organization and Selection in
Evolution. Oxford: Oxford University Press.
Kauffman, S. (1995). At Home in the Universe: The Search for Laws of Self-
Organization and Complexity. Oxford: Oxford University Press.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
7
3

Historical Ecology and the Longue Durée 37

Kidambi, P. (2011). Time, temporality and history. In S. Gunn and L. Faire, eds.,
Research Methods for the Arts and the Humanities: Research Methods for
History, pp. 220–37. Edinburgh, GBR: Edinburgh University Press. ProQuest
ebrary. Web. 2 November 2014.
King, D. (2005). Finding Atlantis: A True Story of Genius, Madness, and an
Extraordinary Quest for a Lost World. New York: Harmony Books.
Knapp, B., ed. (1992). Archaeology, Annales and Ethnohistory. Cambridge:
Cambridge University Press.
Lane, D., Pumain, D., van der Leeuw, S., & West, G., eds. (2009). Complexity
Perspectives in Innovation and Social Change. Springer.
Le Roy Ladurie, E. (1988). Times of Feast, Times of Famine: A History of Climate
since the Year 1000. New York: Noonday Press – Farrar, Straus & Giroux.
Lenton, T. M. (2013). What early warning systems are there for environmental
shocks? Environmental Science & Policy, 27, S60–S75.
Levin, S. A. (1992). The problem of pattern and scale in ecology. Ecology, 73,
1943–67.
Levin, S. A. (1998). Ecosystems and the biosphere as complex adaptive systems.
Ecosystems, 1, 431–6.
Linnaeus, C. (1735). Systema naturae, sive regna tria naturae systematice prop-
osita per classes, ordines, genera, & species. Leiden: Haak.
Lubbock, J. (1865). Pre-historic Times as Illustrated by Ancient Remains, and the
Manners and Customs of Modern Savages. London & Edinburgh: Williams
and Norgate.
Lucas, G. (2005). The Archaeology of Time. London: Routledge.
Lyell, C. (1830). Principles of Geology. 1st edition. London: John Murray.
May, R. M. (1977). Threshold and breakpoints in ecosystems with a multiplicity
of states. Nature, 267(5628), 471–7.
Mbiti, J. S. (1990). African Religions and Philosophy. London: Heinemann.
McGlade, J. (1995). Archaeology and the ecodynamics of human-modified land-
scapes. Antiquity, 69(262), 113–32.
McGlade, J. (1999). The times of history: archaeology, narrative and non linear
causality. In T. Murray, ed., Time and Archaeology, pp. 139–63. London:
Routledge.
McGlade, J. (2006). Ecohistorical regimes and la longue durée: an approach to map-
ping long term societal change. In E. Garnsey and J. McGlade, eds., Complexity
and Co-evolution, Continuity and Change in Socio-economic Systems, pp.
77–114. Cheltenham: Edward Elgar.
Meadows, D. H. (2008). Thinking in Systems: A Primer. White River Junction,
VT: Chelsea Green Publishing.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
8
3

38 Sinclair et al.

Merriam-Webster (2017). Retrieved 24 February. www.merriam-webster.com/


words-at-play/origin-of-the-phrase-tipping-point.
Merrill, R. M. (2012). An Introduction to Epidemiology, 5th edition. Sudbury,
MA: Jones & Bartlett.
Moen, J. & Keskitalo, E. C. H. (2010). Interlocking panarchies in multi-use boreal
forests in Sweden. Ecology and Society, 15(3), 17.
Murray, T. (1999). Time and Archaeology. London: Routledge.
Needham, J. (1954–9). Science and Civilization in China. Cambridge: Cambridge
University Press.
Odum, E. P. & Odum, H. T. (1953). Fundamentals of Ecology. Philadelphia:
W. B. Saunders.
Olsson, L., Jerneck, A., Thoren, H., Persson, J., & O’Byrne, D. (2015). Why resil-
ience is unappealing to social science: theoretical and empirical investigations
of the scientific use of resilience. Science Advantages, 1(4), e1400217.
Pace, M. L., Carpenter, S. R., & Cole, J. J. (2015). With and without warning: man-
aging ecosystems in a changing world. Frontiers in Ecology and Environment,
13, 460–7.
Parrott, L. & Lange, H. (2013). An introduction to complexity science. In C.
Messier, J. Klaus, K. Puettmann, & D. Coates, eds., Managing Forests as
Complex Adaptive Systems: Building Resilience to the Challenge of Global
Change. London: Earthscan, Routledge.
Peterson, G. D. (2002). Contagious disturbance, ecological memory, and the emer-
gence of landscape pattern. Ecosystems, 5, 329–38.
Pielou, E. (1975). Ecological Diversity. New York: Wiley.
Redman, C. L. (2005). Resilience theory in archaeology. American Anthropologist,
107(1), 70–77.
Robinson, K. S. (2009). Galileo’s Dream. London: HarperVoyager.
Rogers, E. (2003). Diffusion of Innovations, 5th edition. New York: Simon &
Schuster.
Samford, P. A. (2007). Pits and the Archaeology of Slavery in Colonial Virginia.
Tuscaloosa: University of Alabama Press.
Scheffer, M., Bascompte, J., Brock, W., Brovkin, V., Carpenter, S., Dakos, V., Held,
H., van Nes, E. H., Rietkerk, M., & Sugihara, G. (2009). Early-warning signals
for critical transitions. Nature, 461(3), 53–9.
Schneider, D. C. (2001). The rise of the concept of scale in ecology. BioScience,
51(7), 545–53.
Scholes, R. J., Reyers, B., Biggs, R., Spierenburg, M. J., & Duriappah, A. (2013). Multi-
scale and cross-scale assessments of social-ecological systems and their ecosys-
tem services. Current Opinion in Environmental Sustainability, 5, 16–25.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
9
3

Historical Ecology and the Longue Durée 39

Scott, J. C. (2009). The Art of Not Being Governed: An Anarchist History of


Upland Southeast Asia. New Haven, CT: Yale University Press.
Soranno, P. A., Cheruvelil, K. S., Bissell, E., Bremigan, M. T., Downing, J. A.,
Fergus, C. E., Filstrup, C., Lottig, N. R., Henry, E. N., Stanley, E. H., Stow,
C. A., Tan, P.-N., Wagner, T., & Webster, K. E. (2014). Cross-scale interac-
tions: quantifying multi-scaled cause-effect relationships in macrosystems.
Frontiers in Ecology and the Environment, 12(1), 65–73.
Spengler, O. (1991) The Decline of the West, vol. 1:21–2. New York: Oxford
University Press (original publication 1918).
Steno, N. (Nicholas Steensen) (1669). De solido intra solidum naturaliter con-
tento dissertationis prodromus. Firenze: Biblioteca Nazionale Centrale.
Sun, Z. Y. & Ren, H. (2011). Ecological memory and its potential applications
in ecology: a review. (in Chinese with English abstract). Chinese Journal of
Applied Ecology, 22(3), 549–55.
Tainter, J. A. (2006). Social complexity and sustainability. Ecological Complexity,
3, 91–103.
Tainter, J. A. & Crumley C. L. (2007). Climate, complexity, and problem solving
in the Roman empire. In R. Costanza, L. J. Graumlich, & W. Steffen, eds.,
Sustainability or Collapse? An Integrated History and Future of People on
Earth, pp. 61–75. Boston: MIT Press.
Tomich, D. (1958). The order of historical time: the longue durée and micro
history. The longue durée and world systems analysis. Histoire et sciences
sociales. Annales E.S.C. XIII, 4.
Trigger, B. (2006). A History of Archaeological Thought, 2nd edition. Oxford:
Cambridge University Press.
Turchin, P. (2003). Historical Dynamics: Why States Rise and Fall. Princeton,
NJ: Princeton University Press.
Van der Leeuw, S., ed. (1998). The Archaeomedes Project – Understanding the
Natural and Anthropogenic Causes of Land Degradation and Desertification
in the Mediterranean. Luxemburg: Office for Official Publications of the
European Union.
Van der Leeuw, S. & McGlade, J. (1997). Archaeology: Time, Process and Structural
Transformations. London: Routledge.
Vance, T. C. & Doel, R. E. (2010). Graphical methods and Cold War scientific practice:
the Stommel diagram’s intriguing journey from the physical to the biological
environmental sciences. Historical Studies in the Natural Sciences, 40(1), 1–47.
Vernadsky, V. (1929). La Biosphère. Paris: Librairie Félix Alcan.
Watson, P. J., LeBlanc, S., & Redman, C. (1971). Explanation in Archaeology: An
Explicitly Scientific Approach. New York: Columbia University Press.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
0
4

40 Sinclair et al.

Webb, C. & Bodin, Ö. (2008). A network perspective on modularity and control of


flow in robust systems. In J. Norberg and G. S. Cumming, eds., Complexity
Theory for a Sustainable Future, pp. 85–118. New York: Columbia
University Press.
Werners, S. E., Pfenniger, S., van Slobbe, E., Haasnoot, M., Kwakkel, J. H. & Swart,
R. J. (2013). Thresholds, tipping and turning points for sustainability under
climate change. Current Opinion in Environmental Sustainability, 5, 334–40.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:26, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.002
1
4

3 Human and Societal


Dimensions of Past Climate
Change
Fredrik Charpentier Ljungqvist

3.1 Introduction
Climate, or rather past climate change, must be regarded as an essen-
tial part of historical ecology, as an interdisciplinary field devoted to
the interactions between environment and humans in the past.* The
direct interaction between historical ecology and palaeoclimatology,
the study of past climate, has so far been surprisingly limited, per-
haps due to academic traditions and disciplinary divisions. There are
some notable exceptions, and there is significant research potential
in the interface between the two fields. It is long known that climate
changes have affected ecological conditions and biophysical systems
across various temporal and spatial scales. Yet it is only recently that
we have gained a more precise understanding of the amplitude of
past climate changes and obtained more detailed and quantitative
palaeoclimatic reconstructions that can be properly compared to, and
analysed with, archaeological and historical data.
The prevailing climate of a region is the most important factor
for its ecosystem, plant growth, and biomass production. The climate
in a region determines whether an area becomes a desert, tropical
rain forest, steppe, tundra, or some other biome, and the climate is
constantly changing in smaller and larger amplitudes and over vary-
ing time scales. Any major change in climate will result in altered
ecological conditions, with shifts in vegetation types and biological
productivity. While weather denotes the short-term (shorter than

* The author is grateful to Dr. Andrea Seim and Dr. Peter Thejll for valuable
comments on parts of this chapter.

41

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
2
4

42 Ljungqvist

ca. thirty years) mean state and variability of the atmosphere with
respect to temperature, precipitation, circulation, air pressure, and
other variables, climate denotes the long-term (longer than ca. thirty
years) mean state and variabilities of these same parameters. Across
wider spatial scales, temperature is the most important, as it drives
changes to many of the other parameters. Generally speaking, from
an ecological perspective, the importance of precipitation tends to
increase with warming climate, and the importance of temperature
tends to increase the colder the climate (for an excellent, albeit in
many respects outdated introduction to climatology, see Lamb 1972–7;
Pierrehumbert 2010).
Various external and internal forcing factors, still inadequately
understood and insufficiently quantified, influence climate change
on regional to global scales (Bradley 1999; IPCC 2013). Over longer
periods of time, larger climatic variations are driven by well-known
and cyclical changes to the Earth’s orbit (tilt and precession) (Berger &
Loutre 1991), the effects of which are amplified by the ocean, sea
ice, and vegetation albedo feedbacks (Wanner et al. 2008; Renssen
et al. 2009). The ‘glacial periods’ or ‘ice ages’ were, for example,
caused by orbital changes. During the Last Glacial Maximum, about
21,000 years ago, these resulted in a global mean temperature depres-
sion of more than 5° C compared to today, and maybe ~15° C lower
annual mean temperature in regions such as Northern Europe, which
had a persistent several-kilometre-thick, high-latitude ice cover. The
extreme cold during the glacial periods reduced the hydrological
cycle and led to much dryer conditions in most of the world. The
glacial periods have dominated the Quaternary period (encompassing
the past 2.6 million years) and their dry conditions caused most of
the non-glaciated part of the world to become steppe or desert
(Andersen & Borns 1997; Alley 2000; Annan & Hargreaves 2013).

3.2 Climate Changes and Their Regional


Expressions during the Holocene
The last glacial period (the Pleistocene) ended approximately
11,700 years ago and the interglacial period since then is known as

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
3
4

Human and Societal Dimensions of Past Climate Change 43

the Holocene. Changes in the Earth’s tilt and precession resulted


in rapid global warming. Maximum warming, the so-called Mid-
Holocene Thermal Maximum (sometimes referred to as the
‘Mid-Holocene Climate Optimum’, ‘Altithermal’, ‘Hypsithermal’,
or ‘Holocene Megathermal’), occurred ca. 8,000–6,000 years ago,
when global mean temperatures exceeded those of today, and in
parts of the higher-latitude areas they were 2–3°C higher (Anderson,
Maasch, & Sandweiss 2007; Karlén & Larsson 2007; Huang et al.
2008; Ljungqvist 2011). During the early and mid-Holocene, the
climate in many places, especially northern Africa and parts of the
Middle East, was much wetter than today. The considerably warmer
conditions of the Northern Hemisphere resulted in a persistent
northward displacement of the Inter-Tropical Convergence Zone
(ITCZ) relative to today. This changed storm track and precipita-
tion patterns and, most importantly, created a northward move-
ment of the Northern Hemisphere monsoon systems (Broecker &
Putnam 2013; Schneider et al. 2014). One obvious effect was that
the Sahara became predominantly steppe and savannah rather than
an extremely dry desert (Tierney et al. 2017). On the other hand,
some of today’s more fertile regions, such as the Great Plains of
North America, experienced considerably drier conditions dur-
ing the same period, due to regional changes in precipitation pat-
terns and, in part, to increased evaporation in the warmer climate
(Sandweiss et al. 1999; deMenocal 2001; Issar & Zohar 2004; Rosen
2007). Some of the main characteristics of the long-term Holocene
climate variability in the Northern Hemisphere are shown in
Figure 3.1.
Beginning ca. 6,000 years ago, the Earth’s climate slowly started
to cool again, initiating a new glacial period caused by changes to
the Earth’s orbit. This long-term cooling is usually referred to as
the ‘Neoglaciation’, named after the renewed glacial expansions.
Although the main (orbital) forcing factor operates slowly, over
millennial time scales, rapid change did occur when the forcing,
or rather the climate response, reached a certain threshold value.
Sudden desertification of the Middle East and the Sahara took place

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
4

44 Ljungqvist

Figure 3.1 Main characteristics of the climate variability in the


Northern Hemisphere from the end of the last glacial period until
present times. A) Schematic graph of millennial-scale variability
of Northern Hemisphere mean temperature during the past 14,000
years (partly adapted from Schönwiese 1995; Ljungqvist 2009). B)
Height of the tree limit in metres relative to AD 1900 (when the tree
limit had reached an exceptionally low position) in the mountains
of northern Scandinavia. The reconstruction is based on pine wood
found at and above the present pine-tree limit (adapted from Karlén
and Kuylenstierna 1996 and updated with Kullman 2015). Given a
temperature lapse rate of 0.7°C per 100 metres of altitude, average early

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
5
4

Human and Societal Dimensions of Past Climate Change 45

around 5,500 years ago (Scheffer et al. 2001; Kuper & Kröpelin 2006;
Holmes 2008; Tierney et al. 2017). The Neoglaciation caused major
reorganisation of the world’s ecological system, regional biomes, and
biomass productivity – albeit of smaller amplitude than during the
transition from Pleistocene to Holocene. The Neoglaciation gave
rise to the major vegetation patterns of the world, many of which
still exist in places where they have not been subsequently altered
by humans.
Superimposed on the millennial-scale orbital forced climate
change, the climate has also shown considerable variability on multi-
centennial to decadal time scales. This variability has partly been
caused by small quasi-cyclical changes in total solar irradiation (Gray
et al. 2010; Steinhilber et al. 2012) and the random occurrence of large
volcanic eruptions (the latter resulting in cooling caused by aerosols
in the stratosphere) (Robock 2000; Sigl et al. 2015). To a large degree,
the climate variability also has been driven by unforced internal
variability of the climate system, with oceanic circulation playing a
major role. Many of the available palaeoclimatic records with suffi-
ciently high resolution and dating control, reflecting changes in tem-
perature, precipitation, or drought, reveal quasi-cyclical periodicities
throughout the Holocene, with the most significant periodicities
lasting around 1,000, 500, 210, and 50–70 years, respectively (Taricco
et al. 2015). At least the length of the 210-year cycle is related to
a well-known periodicity in solar variability, the Suess cycle a.k.a.
the de Vries cycle (concerning climate oscillations on different time

Figure 3.1 (continued)


and mid-Holocene Scandinavian summer temperatures were at the very
least 2°C higher than during the Little Ice Age whereas temperatures
during the Roman Warm period and the Medieval Warm Period seem
to have been at least ca. 1.5°C higher. C) Stable isotope data from a
speleothem record from Dongge Cave, southern China, showing East
Asian summer monsoon intensity during the past 14,000 years by
Dykoski et al. (2005). Note that the summer monsoon generally has
been stronger (and reached further northward) when the climate has
been warmer in the Northern Hemisphere.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
6
4

46 Ljungqvist

scales, see Bond et al. 1997, 1999, 2001; Wagner et al. 2001; Raspopov
et al. 2004, 2008; Willard et al. 2005; Liu et al. 2011; Breitenmoser
et al. 2012).
Even during periods of more or less consistent global warming
or cooling, the amplitude of the temperature increase or decrease has
not been uniform around the globe. As a rule, changes have been the
largest at higher latitudes and smallest in the tropics. This increase
in amplitude of temperature in accordance with latitude is known
as ‘polar amplification’ or ‘Arctic amplification’ (Hind et al. 2016).
There is also a prominent land-sea gradient in the amplitude of tem-
perature variability, with generally much larger variability over land
than over the oceans. While changes in temperature – regardless of
cause – show a rather coherent large-scale pattern, at least in terms of
the direction of change, the effects on the hydroclimate are of a dis-
tinctively localised pattern due to, among other factors, orographic
features, water bodies, vegetation cover and land use (such as irriga-
tion), and, consequently, evaporation. When the climate warms or
cools, it can in many regions change the amount and/or seasonality of
the precipitation, making some regions wetter and others dryer. Parts
of the subtropics, in particular, tend to become drier in a warmer
climate, while wetter conditions usually prevail in much of the tem-
perate zone and part of the equatorial zone (Held & Soden 2006; Cook
et al. 2014). Note, however, that a large body of evidence indicates
that the spatial signature of precipitation change shows considerable
variability from one warm or cold period to the next (this has been
investigated in greatest detail for China where this is clearly the case;
see Hao et al. 2012, 2016; for the whole Northern Hemisphere, see
Ljungqvist et al. 2016).

3.3 The Climate of the Past Two Millennia


In recent years the possibilities of assessing detailed spatial and tem-
poral patterns of climate change for approximately the past two mil-
lennia have improved (see Fact Box 3.1 and Figure 3.2 for examples
of recent temperature reconstructions for Europe and China,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
7
4

Human and Societal Dimensions of Past Climate Change 47

respectively). The generally warm period from ca. BC 300 to AD


300 was in spite of some cold spells, often referred to as the ‘Roman
Warm Period’ (Crumley 1993; Luterbacher et al. 2016). Colder condi-
tions followed in the period between ca. AD 300 and AD 800, usu-
ally referred to as the ‘Dark Age Cold Period’, with the maximum
cooling during the sixth and seventh centuries recently termed the
‘Late Antique Little Ice Age’ (Büntgen et al. 2016). Considerable peri-
ods during the Dark Age Cold Period were wet over extensive parts
of Western Europe, causing frequent flooding and abandonment of
low-lying fields and settlement sites (Cheyette 2008). The seem-
ingly coldest period during the past two millennia in the Northern
Hemisphere, or even globally, followed two large volcanic eruptions
in AD 536 and AD 540 (Stothers 1999; D’Arrigo et al. 2001; Arjava
2005; Larsen et al. 2008; Sigl et al. 2015). This cooling resulted in
severe harvest failures, famine, and large-scale settlement abandon-
ment/depopulation, at least in Northern Europe and northern China
(Gunn 2000; Gräslund & Price 2012; Löwenborg 2012; Widgren 2012;
Tvauri 2014; Toohey et al. 2016).

Fact Box 3.1: Palaeoclimate Proxy Data


Regular meteorological measurements were introduced in
Europe in the eighteenth century and in the rest of the world in
the nineteenth century (Jones 2016). In order to obtain earlier
information about climate variability, indirect (proxy) data must be
used to trace changes in temperature, precipitation, drought, and
circulation patterns. Quantitative estimates of different climate
variables, albeit often with large uncertainties, can be obtained
by palaeoclimatologial approaches, such as applying different
statistical transfer or calibration models to climate proxy data and
instrumental measurements. Many different types of proxy data
are presently used for climate reconstruction purposes and all have
their different strengths and limitations (Bradley 1999; Jones et al.
2009; Christiansen & Ljungqvist 2017). Some of the proxies most
commonly used will now be briefly described.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
8
4

48 Ljungqvist

For the study of the past millennium, documentary records,


such as tax records, harvest yield records, diaries, and gazettes
with information about past weather conditions are very useful.
Such written sources, mainly available for Europe and East Asia,
can in some cases provide rather exact temperature information,
e.g. the date of the freezing/thawing of a body of water. The dates
of agricultural activities, such as sowing and harvesting, may also
provide climate information, but this type of data are usually
less precise. General descriptions of temperature, precipitation,
or drought are often more difficult to transform into quantitative
climate data (Brázdil et al. 2005; Ge et al. 2008; Jones et al. 2009).
The most important branch of palaeoclimatology for the late
Holocene is dendroclimatology, i.e. the study of year-to-year
variations in tree-ring chronologies to track past climate variability
(Fritts 1976; St. George 2014; Esper et al. 2016). Tree-ring records
have absolute annual dating control and ample material can be
obtained from living, historical, archaeological, fossil, and sub-fossil
material. Methods have been developed with the aim of recovering
the low-frequency climate signal exceeding the length of individual
tree segments (Cook et al. 1995; Esper et al. 2003; Melvin & Briffa
2008), making it possible to reconstruct multi-centennial climate
variability from tree-ring data. The only tree-ring records that are
truly useful for climate reconstruction purposes, however, are those
obtained in environments where tree growth is primarily controlled
by a single climate parameter (Fritts 1976; Babst et al. 2013; St.
George 2014; Hellmann et al. 2016). In cold regions, warm season
temperature constitutes the most important parameter, while in
dry areas, the important factor for tree growth is usually the water
availability over a whole year, or a particular season. Climate
signals are most often obtained from tree-ring width (wider rings
representing more favourable climatic growing conditions), but a
far better proxy for temperature is the maximum latewood (greatest
density of wood formed towards the end of the growing season)
density in the tree rings (Esper et al. 2016). Recently, a number of
tree-ring-based annual isotope records have also been developed that
may, besides temperature or precipitation, reflect climate parameters
such as cloudiness.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
9
4

Human and Societal Dimensions of Past Climate Change 49

Glacial ice-core data, in particular stable isotope (δ18O) data, which


in some cases cover hundreds of thousands of years, can be obtained
from polar ice caps and mountain glaciers (Jouzel 2013). At sites
with high snow accumulation, ice-core data can have annual to
decadal resolution for thousands of years. Whereas δ18O is usually
considered to reflect temperature in high-latitude regions, the
seasonality and source area of precipitation can be more important
factors at lower latitudes. Ice cores are moreover very important in
palaeoclimatology since these can provide volcanic aerosol data,
giving us knowledge of when large volcanic eruptions occurred and
their amplitude. From Beryllium-10 (10Be) values in the ice we can
also estimate past total solar irradiance (TSI, Steinhilber et al. 2012),
as well as past levels of greenhouse gases (Bender et al. 1997).
In addition, speleothem records from caves are increasingly used
for studying climate change (Fairchild & Baker 2012). While at
some sites their growth and isotope values may reflect temperature,
speleothems are above all sensitive to the amount and seasonality
of precipitation, especially so in monsoon regions. In particular,
the variability of the East Asian summer monsoon has successfully
been reconstructed with the help of speleothem data throughout the
Holocene (Chen et al. 2016).
For longer time scales, terrestrial and marine sediment archives,
including pollen records, are the most important source of climate
information (Bradley 1999). Such archives are only rarely annually
laminated, but typically have multi-decadal to centennial sample
resolution and are often relatively dated through radiocarbon
dating. However, their strength lies in low-frequency climate
changes at multi-centennial or longer time scales. Marine sediment
reconstructions based on chemical or biological proxies of past sea
surface temperature, sometimes covering hundreds of thousands
of years, are available from many parts of the world’s oceans
(Leduc et al. 2010). Terrestrial sediment archives, especially pollen
data, have been used to reconstruct changes in precipitation and
temperature, often on a seasonal basis, for many parts of the world.
This type of data has been highly important for our knowledge of
climates during the Last Glacial Maximum and the earlier part of the
Holocene (Bartlein et al. 2011).

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
0
5

50 Ljungqvist

(a) Summer temperature


Temperature anomalies in °C 30 year average
relative to 1961–1990 2

–2

1 200 400 600 800 1000 1200 1400 1500 1800 2000

(b)
Annual mean temperature (decadal mean)
Temperature anomalies in °C

0.8
Uncertainty
relative to 1961–1990

–0.8

1 200 400 600 800 1000 1200 1400 1500 1800 2000
Year (AD)

Figure 3.2 Reconstructed temperature variability in Europe and


China, respectively, during the past two millennia. The temperature
for both Europe and China is shown as anomalies in °C relative to
1961–90. A) Mean European summer (June–August) temperature from
138 BC to 2003 AD reconstructed from temperature-sensitive tree-
ring data and historical documentary data by Luterbacher et al. (2016).
For simplicity’s sake, the uncertainty estimates and the instrumental
temperature data from 1850 to the present are not shown in the graph.
B) Decadal-scale annual mean temperature variations in China during
the past 2,000 years reconstructed from various types of temperature-
sensitive palaeoclimate proxy data by Ge et al. (2013) with uncertainty
estimates shown in grey.

A longer period with warmer climate conditions prevailed


again during the ‘Medieval Warm Period’, or the ‘Medieval Climate
Anomaly’, which lasted between ca. AD 800 and AD 1250. In most
locations, the warming peaked during the second half of the tenth cen-
tury, at least in the Northern Hemisphere, and for a couple of decades
the warming may have equalled – and regionally even exceeded – the

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
1
5

Human and Societal Dimensions of Past Climate Change 51

magnitude of the post-1990 global warming. Relatively warm condi-


tions also seem to have been rather common in the last part of the
twelfth and the first half of the thirteenth century (Mann et al. 2008,
2009; Christiansen & Ljungqvist 2012; Ljungqvist et al. 2012; Wilson
et al. 2016). The hydrological changes across much of the world were,
arguably, of greater importance than the higher temperatures with
the ‘megadroughts’ in central and western North America and in
Central America as perhaps the most prominent examples (for medi-
eval megadroughts in North America, see in particular Hughes &
Graumlich 1996; Laird et al. 1996; Benson et al. 2002; Cook et al.
2004; Cook et al. 2014; see also Figure 3.3). Wetter conditions pre-
vailed in other regions, notably in central and north-eastern China
and large parts of India, mainly attributed to enhanced monsoon
precipitation (see Figure 3.4). This has been caused by a northward
movement of the Inter-Tropical Convergence Zone, as a result of the
warming of the Northern Hemisphere (Yang et al. 2014).
The transition towards colder conditions began in the late thir-
teenth and early fourteenth centuries – possibly somewhat later in
the Southern Hemisphere – and continued until the late nineteenth or
even the early twentieth century. This global cooling is known as the
‘Little Ice Age’ (ca. AD 1300–1850). The coldest conditions occurred
in most regions during the seventeenth century, but the mid-fifteenth
and the early eighteenth centuries were also very cold (Moberg et al.
2005; Ljungqvist et al. 2012; PAGES 2k Consortium 2013). The Little
Ice Age was in fact, in many locations, probably the coldest interval
of such length since the end of the last Ice Age (Mayewski et al. 2004;
Matthews & Briffa 2005; Wanner et al. 2008, 2011). Mountain gla-
cier expansion was noticeable from the tropics to the polar regions
(Solomina et al. 2016); areas with permafrost expanded and altitu-
dal and latitudal tree lines retreated in many places (Payette et al.
1989; Hiller et al. 2001; Mazepa 2005; Kullman 2015). Nevertheless,
the Little Ice Age was not uniformly cold. Especially the early six-
teenth and the mid-eighteenth centuries experienced quite mild, and
regionally even rather warm conditions (see Ljungqvist et al. 2012).

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
2
5

52 Ljungqvist

100
Percentage drought area in western DRIER
80
Mean medieval drought level
(AD 900–1300)
60 WETTER
United States

40

20
Mean 20th century
0 drought level

800 900 1000 1100 1200 1300 1400 1500 1600 1700 1800 1900 2000
Year (AD)

Figure 3.3 Reconstructed long-term aridity changes (AD 800–2003)


in the western half of the United States derived from the tree-ring
width–based North American Drought Atlas by Cook et al. (2004,
2007). The graph shows the percentage of the land area in the western
half of the United States affected by drought (as defined by a value
of –1 or lower in the Palmer Drought Severity Index). Annual values
are shown as thin grey lines and sixty-year averages as the thick black
line. The average drought in the early 2000s is shown with a black
asterisk. The North American Drought Atlas is based on hundreds
of long, moisture-sensitive, tree-ring-width chronologies, which
allows for the reconstruction of changes in relative drought in time
and space from annual to multi-centennial time scales. The average
drought area during medieval times (AD 900–1300) and during the
twentieth century, respectively, is indicated by thick grey lines. Note
that the twentieth century has overall been significantly wetter than
the long-term average, and that the early twentieth century (until ca.
1930) in the western half of the United States was one of the wettest
periods of the whole past twelve centuries. When allocating the water
resources, and planning the water supply, in this arid and semi-arid
region the exceptionally wet early twentieth century has in the past,
unfortunately, to a large degree been used as a benchmark. This
highlights the need to include a long-term palaeoclimate perspective in
water planning to account for natural long-term variations in drought
and precipitation. Anthropogenic global warming is expected to make
large parts of the southwestern United States even drier in the future
than during the medieval ‘megadroughts’, although these climate
model-based projections have large uncertainties.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
3
5

Human and Societal Dimensions of Past Climate Change 53

(a) Stronger moonson


(wetter)

Weaker moonson
(drier)

200 400 600 800 1000 1200 1400 1600 1800 2000

(b) Stronger moonson


(wetter)

Weaker moonson
(drier)

600 800 1000 1200 1400 1600 1800 2000


Year (AD)

Figure 3.4 Reconstructed relative strength of the summer monsoon


in central China and northeast India, respectively, during the past
one to two millennia from stable isotope cave speleothem data. A)
Reconstructed relative monsoon strength in central China from stable
isotope data in a speleothem record from Wanxiang Cave near the
present-day northern limit for the East Asian summer monsoon by
Zhang et al. (2008). B) Reconstructed relative Indian summer monsoon
strength from stable isotope data in speleothem records from Jhumar
and Wah Shikar Caves in northeast India by Sinha et al. (2011). Multi-
decadal periods of weak monsoons coincide with frequent and severe
harvest failures, resulting in widespread famines. Note that the
weakening of the Indian summer monsoon occurred earlier during the
Little Ice Age than the weakening of the East Asian summer monsoon.

This cold period affected precipitation patterns in some places.


The monsoon circulation was affected by the displacement of the
Inter-Tropical Convergence Zone closer to the Equator (Schneider
et al. 2014). Generally wetter conditions prevailed in most of North
America and in the western Mediterranean, while parts of East Asia

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
4
5

54 Ljungqvist

(e.g. China and India) experienced an increased number of monsoon


failures and droughts (Cook et al. 2004, 2010; Zhang et al. 2008; Chen
et al. 2015).
From ca. 1920 to 1950, the global mean temperature increased
considerably, primarily in the extra-tropical Northern Hemisphere
(Jones 2016). This first major phase of this modern warming was, to a
large extent, influenced by natural forcing factors, such as increased
solar irradiance, internal variability, and a lack of major volcanic erup-
tions. Unlike the warming of the late twentieth century, the early
twentieth-century warming was mainly restricted to the mid and
high latitudes of the North Atlantic and nearby regions. The signature
of the warming was a lot weaker, or almost absent, in other regions.
A temperature stagnation, and in many areas actually a temperature
reduction, occurred from ca. 1950 to 1975. Subsequently, strong glo-
bal warming, evident in almost all regions, took place between ca.
1975 and 2000 (Hansen et al. 2006, 2010; Compo et al. 2011; Morice
et al. 2012; Rohde et al. 2013; Jones 2016). The main driving force
behind this was the increase in anthropogenic greenhouse gas emis-
sions (mainly carbon dioxide, CO2) from the burning of fossil fuels
(IPCC 2013 and references therein). The strong warming trend has
continued in the Arctic, but has stagnated on a global level, with
rather stable global mean temperatures between ca. 2000 and 2013,
even showing a slightly decreasing trend in parts of the mid latitudes
and the tropics (Gleisner et al. 2015). This ‘global warming hiatus’
was most likely a result of natural climate variability, and perhaps
decreased total solar irradiance. However, significant global warming
is predicted to continue in the next 100 years (IPCC 2013). Indeed,
the years 2014 and 2015 saw renewed warming, arguably ending the
hiatus, and the warming continued in 2016, resulting in the highest
global mean temperatures since a global network of instrumental
meteorological measurements started in the mid-nineteenth century
(Jones 2016). Nevertheless, the rate and magnitude of global warming
continues to be somewhat lower than most climate model simula-
tions predict (Fyfe et al. 2016).

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
5

Human and Societal Dimensions of Past Climate Change 55

The continuation and increase of anthropogenic greenhouse


gas emissions into the atmosphere will force the climate to become
warmer, resulting, within perhaps only a couple of decades, in tem-
peratures exceeding those of the mid-Holocene Thermal Maximum.
More significantly, future global warming is expected to alter pre-
cipitation patterns with, in places, severe consequences for the eco-
system. The magnitude of future global warming is highly uncertain,
but the best estimate, provided by the Fifth Assessment Report (AR5)
of the Intergovernmental Panel on Climate Change (IPCC 2013), is
that by AD 2100, the global warming is most likely to exceed 1.5°C
and may exceed 4.5°C, relative to 1850–1900.

3.4 Impacts of Climate Change on Human


Well- Being in Past Societies
Palaeoclimatological research has made increasingly clear that the
previously predominant view of a rather stable climate during the
Holocene – or at least during the past two millennia – is simply not
correct. Archaeologists and historians therefore need to re-evaluate
the historic role of climate change on societies, and in some cases
even its potential impact on the rise and fall of civilisations (for
recent discussion, see d’Alpoim Guedes et al. 2016; McNeill 2016).
Although small compared to the Pleistocene, the Holocene cli-
mate fluctuations and their amplitude may have had, especially on
regional scales, a significant impact – positive or negative – on the
agricultural carrying capacity and resilience of a number of socie-
ties. Changing climate conditions may also have triggered large-scale
migration. A particular climate change (e.g. a warming or cooling)
usually had different regional impacts, not least by increasing pre-
cipitation amounts in some regions and decreasing ones in others,
thus resulting in larger agricultural productivity in some places but
harvest failures in others (see Parry et al. 1988a, 1988b for examples
of impact on agricultural productivity by climate changes).
Although debated, it has been suggested that climate change
has had a significant impact on a number of past societies; some even

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
6
5

56 Ljungqvist

seem to have collapsed under the pressure of adverse climate change


when other unfavourable external and/or internal non-climatic fac-
tors were present. Prominent examples are arguably the collapses,
induced by drought, of the Akkadian Empire ca. 2100 BC (Cullen
et al. 2000; Issar & Zohar 2004; Finné et al. 2011) and the Classic
Maya civilisation ca. AD 900 (Gill 2000; Lucero 2006; Douglas et al.
2016). The late medieval demise of Norse Greenland seems, on the
other hand, to have been caused by a cooling climate (Buckland et al.
1996; Barlow et al. 1997; Dugmore et al. 2007; Madsen 2014). It has
also been established that major dynastic collapses and transitions in
China frequently occurred during regionally dry periods (Fang & Liu
1992; Chu & Lee 1994; Zhang et al. 2007, 2010; Su et al. 2016). Cold
and/or dry periods in Chinese history also exhibit significantly higher
proportions of population decline, nomad invasions, rebellions, and
wars (D. D. Zhang et al. 2007; Z. Zhang et al. 2010). Although still a
matter of controversy, it has also been suggested that the population
movements and armed invasions of Europe from Central Asia during
the Migration Period, which coincide with the Dark Age Cold Period,
were at least partly caused by less favourable climate conditions on
the steppes (McCormick et al. 2012; Büntgen et al. 2016). Restricting
ourselves to the much-studied impacts of the Medieval Warm Period,
it has been demonstrated that the warming helped facilitate a major
expansion of agriculture and population in Northern Europe and the
northern parts of East Asia (Lamb 1977; Ljungqvist 2009). Yet, in
other regions, most notably North America, this resulted in severe
droughts due to changed precipitation patterns. This caused societal
crises among, e.g. the Classic Maya civilisation of Central America
(Gill 2000; Lucero 2006; Douglas et al. 2016) and the Anasazi cul-
ture in the southwestern parts of the United States (Cook et al. 2004;
Benson et al. 2007; Blinman 2008; Benson & Berry 2009; Benson
2010; Bocinsky & Kohler 2014).
Increasingly, archaeologists researching ancient civilisa-
tions acknowledge that climate change can have significant socio-
economic impact. Historians mainly studying societies from the

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
7
5

Human and Societal Dimensions of Past Climate Change 57

past two millennia have, however, been generally more reluctant.


They generally acknowledge that year-on-year weather variations
and extreme weather events have had considerable impact on the
well-being of premodern societies. In terms of long-term climate
change, such as the transition from the Medieval Warm Period to
the Little Ice Age, historians often see these as too small and insig-
nificant to have had any noticeable socio-economic or demographic
impact, with the exception of a few marginal areas (e.g. Iceland and
Greenland). This is despite increasing palaeoclimatic evidence point-
ing to the opposite (for updated and critical reviews of historians’
approaches to climate change, see Carey 2012; Parker 2013; Brooke
2014; Fan 2015; Campbell 2016; McNeill 2016; Slavin 2016).
Climate change has, admittedly, rarely been the only or even
the primary cause of larger societal change, but can arguably be con-
sidered as an important contributing factor in a number of cases. One
reason why historians and, to a lesser extent, archaeologists neglect
the importance of climate change is probably the legacy of the over-
simplified ‘environmental determinism’ prevailing in the first half
of the twentieth century (the most famous example is Huntington
1907; see also Huntington 1913). This is, however, hardly the only
reason; indifference to the impact of climate change on past societies
can also be attributed to the tendency among scholars within the
humanities to diminish the importance of external factors for soci-
etal change, in favour of internal, cultural factors. The main reason,
however, is presumably that historians and archaeologists usually
have limited knowledge of the rapid advances in the modern field
of palaeoclimatology. Until very recently, palaeoclimatologists were
usually unable to deliver quantitative climate reconstructions that
were detailed and reliable enough to be applicable and useful to his-
torical studies.
A number of studies, mainly by American and Chinese schol-
ars, have shown that, at the macro level over longer periods of time,
there are significant correlations between climate change and demo-
graphic development and the intensity of warfare in China and Europe

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
8
5

58 Ljungqvist

(for updated reviews of the impact of climate change on the history


of the past two millennia, see Behringer 2010; Zhang et al. 2011;
Carey 2012; McMichael 2012; Parker 2013; Brooke 2014; Diaz &
Trouet 2014; Hsiang & Burke 2014; Fan 2015; Campbell 2016; Nelson
et al. 2016; more popular syntheses are Fagan 2000; Diamond 2005;
Glaser 2008; Ljungqvist 2009). In particular, a highly important and
statistically verified relationship between temperature and popula-
tion growth rate has been observed across geographic regions from
China to Europe for the late medieval and early modern periods. The
cold spells of the Little Ice Age, interestingly, show lower population
growth rates, even in the warmer parts of the world. Even more note-
worthy, nearly 70 per cent of the documented population collapses
in the world between AD 800 and AD 1900 occurred during the few
short periods of sharp cooling, which in total comprise less than
2 per cent of this period (Lee & Zhang 2015). In some of the dry and
warm regions of the world, however, population collapse commonly
occurred not only during the Little Ice Age, but even more frequently
during the Medieval Warm Period (for a synthesis, see Lee & Zhang
2015 and cited literature therein). Recent quantitative studies of the
impact of climate change on humans in past societies at an aggre-
gated level (Figure 3.5) have thus proven that such a relationship does
indeed exist (see e.g. Fan 2010; Zhang et al. 2011; Lee & Zhang 2015
and references therein). Importantly, these studies have mainly used
older palaeoclimatological data showing smaller and less constrained
large-scale changes than those supported by the most recent palaeo-
climatic reconstructions. The actual effects of climate on human
well-being over broader spatial scales may thus be even larger and
more significant than the studies by, e.g. Zhang and colleagues (2011)
and Lee and Zhang (2015) support.

3.5 Causal Climate– Human Relationships


and Societal Responses
Statistical studies conducted at macro levels through the correla-
tion of palaeoclimate time series with different historical indices of

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
9
5

Human and Societal Dimensions of Past Climate Change 59

Climate Change
temperature, precipitation, drought

Agricultural productivity
harvest yield

Food supply per capita


availability and price of food

Social unrest Migration Food shortage


lawlessness “climate refugees” malnutrition, famine

Armed conflicts Epidemics Poor nutritional status


rebellions, war dysentery, plague poor health

Population
Population growth rate
and population size

Figure 3.5 Schematic overview of causal linkages between climate


changes with adverse changes for agricultural productivity and human/
societal crises in the premodern world (partly inspired by figure 14.7
in Lee & Zhang 2015). The terms typeset in bold face define various
factors linking adverse climate changes to impacts on human
populations and societal stability, and the terms typeset in italics
represent examples of these factors. Boxes shadowed in grey represent
the most important causal climate–societal linking factors. The black
lines with arrows shows the directions of the chain of impacts. Thick
black lines show the generally most important causal climate–societal
links whereas thin black lines show the weaker, albeit still generally
significant, causal climate–societal links.

human well-being – although important – do not really demonstrate


the causal relationships that link climate change and human history.
Which are the mechanisms that have led to climate change either
increasing or decreasing food security and societal stability? We may
begin by stressing that the impact of climate change has been – and
arguably still is – most significant near borders or transition zones
between different climate and vegetation zones. Regions near ther-
mal or hydrological limits of certain types of agricultural or pastoral

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
0
6

60 Ljungqvist

activities are the most sensitive to climate change, since warmer/


colder or dryer/wetter conditions displace such borders. From a
human perspective, the largest climate-change-related gains or losses
are made in such marginal locations. In regions where certain types
of agriculture or pastoral activities are already under climatic strain,
in a worst-case scenario, they may become almost impossible to con-
duct in a colder or dryer climate. Conversely, climate change may
push the borders for certain types of agriculture or pastoral activity
into new, previously climatically unsuitable regions.
Already in prehistory, climate change affected the potential
for human subsistence in very direct ways, as this altered the eco-
system or biomass productivity, influencing the availability of game
or triggering wild animal migration. The hunter-gatherers then had
no choice but to follow or to move to better hunting grounds. With
the introduction of agriculture and permanent settlements, this flex-
ible form of adaptation in line with ecological constraints of climate
change almost disappeared. Since the populations of most premod-
ern societies lived with small margins and limited food surpluses,
even modest climate-driven effects on agricultural yields could have
a large impact on food supply. Famines may occur, depending on the
availability of alternative food sources, even with a relatively modest
reduction of the yield and especially if they occurred in consecutive
years (see e.g. Parry 1978; Abel 1980; Parry & Carter 1983; Alexander
1987; Ljungqvist 2009; Parker 2013; Campbell 2016).
The degree of agricultural marginality can be conceptualised
as the risk or frequency of harvest failure. Climate change could affect
the risk of harvest failures in a given region, although this may vary
among crops. Consecutive years of bad harvests or outright harvest
failures were particularly problematic for food security in premod-
ern societies and could easily cause major subsistence crises. This
was more likely to occur in marginal regions with specific climates
and occasionally triggered subsistence crises affecting entire socie-
ties. Reduced agricultural productivity resulting in famine could lead
to demographic change, especially among the poorer strata of the

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
1
6

Human and Societal Dimensions of Past Climate Change 61

population, by increasing mortality and reducing fertility rates. Poor


nutrition increased vulnerability to disease and thus contributed sig-
nificantly to epidemics and mortality.
Famine could also have major and even devastating effects for
elite members of society, even if they were rarely directly affected.
The reduction in agricultural productivity could be accompanied
by a stagnation or decrease in population, tax revenue, land rent,
and other income necessary for maintaining elite status and state
functions. Famine reduced the capacity to construct monumen-
tal buildings and keep large armies, servants, and other employees,
thus potentially resulting in unemployment, furthering poverty and
triggering social unrest. Moreover, when secular or religious leaders
proved unable to deliver relief during famine, their popularity and
authority most likely decreased (see the discussion in Sorokin 1975;
Bryson 1977; Rotberg & Rabb 1985; Arnold 1988; Halstead & O’Shea
1989; Hugo & Currey 1989; Newman 1990). As a general rule, any
attempt to claim the same revenues as before from the population, in
spite of reduced agricultural productivity and dearth, or to increase
the resource extraction per capita after a crisis to compensate for pop-
ulation/productivity loss, could lead to social unrest, open rebellion,
or even civil war.
The reactions to the adverse effects of climate change and the
choice of adaptation strategies – if any – have varied from society
to society. For those affected, climate change was not always obvi-
ous or easily distinguished from ordinary year-to-year variability. It
could develop gradually, e.g. through an increase in warm years or
a decrease in cold years, and only become evident after a few dec-
ades. The full range of adaptation options to mitigate adverse climate
change has rarely been utilised. Different societies have been better
or poorer at coping with or, indeed, taking advantage of the effects of
climate change.
In some sense, when studying the human and social dimensions
of climate change, it is more relevant to focus on how the impact
has been handled than on the climate change itself (Figure 3.6). This

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
2
6

62 Ljungqvist

Adverse climate change or extreme weather event

Biophysical focus
First order impact: biophysical effects
Primary production: Diverse impacts on planted
reduced biomass crops with different
altitude and exposure etc.

Second order impact: human and societal well-being


Prices of food, Market systems,

Socio-economic focus
epidemics and transport infrastructure
epizootics

Third order impact: demographic and social implications


Malnutrition, demographic Food supply, migration,
growth (mortality, fertility), public and private welfare,
social conflicts societal resilience to crises

Cultural focus
Fourth order impact: cultural responses
Crisis interpretation, Religious rituals, amendments
cultural memory, (e.g. marriage law),
learning processes adjustments, adaptations

Figure 3.6 The different orders of impact from adverse climate change
or adverse extreme weather events ranging from the direct biophysical
effects on plant growth down to cultural responses (adapted in revised
form from Krämer 2015; Luterbacher & Pfister 2015). This schematic
view shows the way adverse climate change or adverse extreme weather
events have affected premodern societies, directly and indirectly, in
a variety of ways (black arrows), and the adjustment and adaptation
strategies that may be attempted to mitigate the effects (grey arrows).
The right axis shows the different foci that the different orders of
impact may typically be analysed from.

observation becomes even clearer when considering that in some


societies, rather modest changes made a considerable impact, while
much larger changes had only negligible impact on others. In this
chapter, the focus is mainly placed on changes with adverse effects
on the agricultural carrying capacity. The reason is that these are the
best documented and, arguably, have the largest effects on society.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
3
6

Human and Societal Dimensions of Past Climate Change 63

Although generalisations are always difficult and sometimes unwar-


ranted, it does seem that societies with a wide range of subsistence
strategies have shown the greatest resilience to climate change and
other environmental impacts. Such societies had ‘safety nets’ when
the ordinary means of subsistence were weakened.
The social organisation and policies for distributing and shar-
ing available food resources in subsistence crises have arguably
been imperative for mitigating the socio-economic consequences
of harvest failures and famine-related mortality in premodern soci-
eties. One illuminating example is the difference in distribution
and affordability of available resources in times of dearth between
early modern England and France. In England, the last major fam-
ine with widespread and high mortality among the poor occurred
in 1601, while in France, famine mortality was recurrent until the
mid-eighteenth century, in spite of the country’s generally more
favourable agricultural conditions (the famine in England of 1741
was restricted to certain regions and sufficiently mitigated in other
regions; see e.g. Post 1985). The main difference was that the poor in
England were guaranteed food through a public welfare system man-
dated by the English Poor Law of 1602, whereas French measures
were a lot more modest and inefficient (Appleby 1980; see also Post
1985 and Walter 1989).
Climate change with adverse effects on agricultural productiv-
ity has at times in some societies been regarded as divine punishment
and therefore mainly addressed through religious measures. The need
for scapegoats may have been rather common, e.g. in Europe, where
during times of dearth in the fifteenth to the seventeenth centuries,
people (predominantly women) were accused of deliberately, through
witchcraft, causing bad weather in order to ruin crops. Although such
‘witches’ were blamed for many different types of ‘crime’, the use
of black magic to ruin crops was, in fact, one of the most common
accusations (see, in particular, Behringer 1995, 1999; Pfister 2007).
Recently, it has also been demonstrated that the persecution of Jews
in Europe became especially common in the period prior to ca. 1600,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
4
6

64 Ljungqvist

when cold/wet periods were causing harvest failure (Anderson et al.


2017), although the causal links still need to be established.
Famine induced by climate change has occasionally resulted
in forced migrations or at least triggered population movements to
more favourable locations. Climate change has also, in some cases,
triggered migrations in order for people to take advantage of oppor-
tunities offered by improved climate conditions in other regions.
Nomadic or semi-nomadic migrations at times covered rather large
distances. Migrations of farming populations were generally shorter,
except when climate-induced famines contributed to trans-oceanic
migration. In the context of climate-change-related migration (argu-
ably in the past, present, and future), it is important to consider the
constraints of many societies on free migration. Political borders and
property rights (e.g. of agricultural land) may limit people’s oppor-
tunities. Infringement of such limitations by migrating populations
easily triggers conflict with those already claiming a territory and its
resources (this topic has been discussed, e.g. by Hsu 1998; Barnett
2003; Afifi & Jäger 2010; Scheffran et al. 2012; Chen 2015; Zhang
et al. 2015 – see also Fact Box 3.2).
In general, ecologically sustainable use of natural resources
appears to be important to successfully mitigate the adverse effects
of climate change on agricultural productivity and its impact on
human and societal well-being. Environmental maladaptation or
overexploitation of natural resources seems to have been a contrib-
uting factor for those cases when adverse climate change has con-
tributed to societal crises or increased the marginality of an already
marginal area (see e.g. Lee 2014, and more generally Ljungqvist
2009). Overpopulation, in relation to the available resources under
given technological conditions, also seems crucial for triggering cri-
ses. Climate change which influences a region’s carrying capacity
negatively may result in overpopulation without an actual popula-
tion increase. Conversely, climate change with positive effects on a
region’s carrying capacity may lower the population pressure on this
region. Negative socio-economic effects of adverse climate change

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
5
6

Human and Societal Dimensions of Past Climate Change 65

Fact Box 3.2: Potential Links between Armed


Conflicts and Climate Change
In the context of ongoing and future global warming, there is growing
interest in possible relationships between climate change and armed
conflict, as well as other types of violence. This concern may partly
be associated with recent outbreaks of war in some dry and warm
parts of the world, although it appears somewhat peculiar, given the
steady reduction in the number and size of armed conflicts, and the
reduced number of casualties in the past decades (for an updated
review of the number and severity of armed conflicts, see Pettersson
and Wallensteen 2015; Melander et al. 2016). Studies of a potential
relationship between armed conflicts and climate change have
above all focussed on regions and periods where climate change has
resulted in reduced agricultural productivity and reduced overall
prosperity (for an updated review, with an extensive bibliography,
see Hsiang et al. 2013; Hsiang & Burke 2014). Studies concerning
both modern conditions and historic societies have usually relied on
statistical correlation between different time series, but contain little
or no actual assessment of causal mechanisms. Consequently, they
have seldom been able to prove and explain the nature of a causal
relationship between climate and conflict, nor have they adequately
assessed alternative explanations for the change in number of
conflicts. The large-scale nature of most such studies has in fact
impeded in-depth understanding of cultural and social conditions
in the different societies involved. Knowledge of these conditions
is necessary for assessing the potential influence of climate change
on conflict (for recent criticism of studies concerning relationships
between climate change and armed conflicts, see in particular
Buhaug 2010; Klomp & Bulte 2013; Meierding 2013; O’Loughlin et al.
2014; Raleigh et al. 2014).
Recent comprehensive syntheses provide some evidence for
a significant increase in the number of conflicts, although not
necessarily their size, in periods when the climate deviates from
its long-term mean. Studies limited to the twentieth century seem
to indicate that higher temperatures have a stronger effect than

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
6

66 Ljungqvist

changes in precipitation/drought, although it is highly questionable


if this result holds true when longer time scales are considered.
Crucially, no causal links have been established explaining why
higher temperatures would result in increased armed conflicts.
Even if recent research supports some form of connection between
climate change and armed conflict and other types of violence at an
aggregated level, over different time scales, it must be emphasised,
this relationship is statistically weak. These studies are, moreover,
associated with numerous problems of a more methodological
character, which raises questions about the validity of the results.
The critical question remains: what causal links could there be
between climate change and armed conflict? The most favoured
explanations seem to relate more to rising criminality and small-scale
violence than large-scale armed conflict. It has been suggested that
when climate change has adverse effects on living conditions and
reduces economic opportunities, violent means of gaining a livelihood
become more attractive. Others argue that economic decline caused
by climate change may result in a collapse of state functions, which
indirectly triggers conflict through lack of means to suppress violence,
thus opening for direct competition over power and resources by use of
force. Still others suggest that social inequality, aggravated by adverse
climate change, is a main cause of conflict. But this last, sometimes
Marxist-inspired explanation seems more applicable to food riots
and revolts than to large-scale armed conflict (e.g. Meierding 2013).
Thus, a truly satisfactory causal link between climate change and an
increasing number of armed conflicts has yet to be established, and in
today’s world, ethnic and religious conflicts are clearly the dominating
cause of armed conflicts. However, as discussed previously in this
chapter, in some cases, climate-triggered migration seems indeed to
have caused conflict and armed competition for resources, in particular
as nomads invaded agricultural areas (Fang & Liu 1992; Chen 2015;
Su et al. 2016). It is therefore possible to conclude this topic by
considering that the historical data do support the idea that an influx
of ‘climate refugees’ into an area can cause social unrest as well as
outright hostility, and that such migrations may at times have taken
the form of armed invasions.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
7
6

Human and Societal Dimensions of Past Climate Change 67

also appear to be related to the lack of buffer zones or empty areas to


expand into, social inflexibility, and an overall inability to adapt to
changing environmental conditions.

3.6 Concluding Remarks


New palaeoclimatological research reveals larger and more sub-
stantial climate changes of the past few thousand years than previ-
ously known, on various temporal and spatial scales. Increasingly,
it is understood that the global climate is a complex and coupled
system where, e.g. cooling of the Northern Hemisphere results in
a southward displacement of the Inter-Tropical Convergence Zone,
leading to a weakening and contraction of the monsoon precipita-
tion systems. The changes in temperature, precipitation, drought,
and other climate variables have, especially on regional scales, been
large enough to have pronounced effects on ecological conditions and
biophysical systems.
Thus climate changes have had substantial effects on the car-
rying capacity for different agricultural and pastoral activities in past
societies. Climate can therefore be considered a key component of
historical ecology. Although climate changes usually either increase
or decrease the annual mean productivity of the prevailing type of
agricultural and/or pastoral farming systems in a given area, it is
typically the changes in the frequency of poor or favourable growing
conditions that have a more pronounced effect. Concerning adverse
climate changes, it is clear that the increased frequency of unfa-
vourable growing conditions and thus harvest failures – especially
the increasing frequency of weak harvests several years in a row –
has resulted in subsistence crises, rather than decreased mean
productivity.
Effects of climate change on human and societal well-being –
regardless of whether the climate change in question has increased
or decreased the potential carrying capacity – must be understood
and analysed within a framework of prevailing social institutions and
human activity. Societal ability to cope with the effects of adverse

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
8
6

68 Ljungqvist

climate changes seems to have been even more dependent on a


society’s social institutions as well as on what is sometimes termed
the social fabric: trust, empathy, creativity, the value of the indi-
vidual, and reliance on the community. These cultural and social fac-
tors highlight the complex and continuously varying relationships
between human societies and environmental changes or constraints
they encounter.
The choice, as well as the success, of mitigation strategies to
handle adverse effects of climate change has been more dependent
on socio-political, socio-economic, and socio-cultural factors than on
the magnitude of the climate change. The study of historical soci-
eties experiencing adverse effects of climate changes reveals some
general tendencies that are at least partly valid for the present and
the future. Factors that have led to reduced resilience include: small
margins due to overpopulation (in relation to resources and technical
constraints), prevailing and severe poverty among large parts of the
population, unequal access to resources, and underdeveloped infra-
structure for large-scale trade, as well as lack of agricultural and eco-
nomic diversity.

References
Abel, W. (1980). Agricultural Fluctuations in Europe: From the Thirteenth to the
Twentieth Centuries. London: Methuen.
Afifi, T. & Jäger, J. eds. (2010). Environment, Forced Migration and Social
Vulnerability. Heidelberg: Springer-Verlag Berlin.
Alexander, P. (1987). Le Climate en Europe au moyen âge: Contribution à l’histoire
des variations climatiques de 1000 à 1425, d’après les sources narratives de
l’Europe occidentale. Paris: Éditions de l’École des hautes études en sciences
sociales [in French].
Alley, R. B. (2000). The Younger Dryas cold interval as viewed from central
Greenland. Quaternary Science Reviews, 19, 213–26.
Andersen, B. G. & Borns, H. W. (1997). The Ice Age World: An Introduction to
Quaternary History and Research with Emphasis on North America and
Northern Europe during the Last 2.5 Million Years. Oslo: Scandinavian
University Press.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
9
6

Human and Societal Dimensions of Past Climate Change 69

Anderson, D. G., Maasch, K., & Sandweiss, D. H., eds. (2007). Climate Change
and Cultural Dynamics: A Global Perspective on Mid-Holocene Transitions.
London: Academic Press.
Anderson, R. W., Johnson, N., & Koyama, D. M. (2017). Jewish persecutions and
weather shocks: 1100–1800. The Economic Journal, 127, 924–58.
Annan, J. D. & Hargreaves, J. C. (2013). A new global reconstruction of tem-
perature changes at the Last Glacial Maximum. Climate of the Past, 9,
367–76.
Appleby, A. B. (1980). Epidemics and famine in the Little Ice Age. The Journal of
Interdisciplinary History, 10, 643–63.
Arjava, A. (2005). The mystery cloud of 536 CE in the Mediterranean sources.
Dumbarton Oaks Papers, 59, 73–94.
Arnold, D. (1988). Famine: Social Crisis and Historical Change. Oxford: Basil
Blackwell.
Babst, F., Poulter, B. & Trouet, V., Tan, K., Neuwirth, B., Wilson, R., Carrer, M.,
Grabner, M., Tegel, W., Levanic, T., Panayotov, M., Urbinati, C., Bouriaud,
O., Ciais, P., & Frank, D. (2013). Site- and species-specific responses of for-
est growth to climate across the European continent. Global Ecology and
Biogeography, 22, 706–17.
Barlow, L. K., Sadler, J. P. & Ogilvie, A. E. J., Buckland, P. C., Amorosi, T.,
Ingimundarson, J. H., Skidmore, P., Dugmore, A., & McGowan, T. H. (1997).
Interdisciplinary investigations of the end of the Norse Western Settlement in
Greenland. The Holocene, 7, 489–99.
Barnett, J. (2003). Security and climate change. Global Environmental Change,
13, 7–17.
Bartlein, P. J., Harrison, S. P. & Brewer, S., Connor, S., Davis, B. A. S., Gajewski,
K., Joel, G., Harrison-Prentice, T. I., Henderson, A. P., Peyron, O., Prentice, I.
C., Scholze, M., Seppä, H., Shuman, B., Sugita, S., Thompson, R., Viau, A. E.,
Williams, J. W., & Wu, H. (2011). Pollen-based continental climate reconstruc-
tions at 6 and 21 ka: a global synthesis. Climate Dynamics, 37, 775–802.
Behringer, W. (1995). Weather, hunger and fear: the origins of the European witch
persecution in climate, society and mentality. German History, 13, 1–27.
Behringer, W. (1999). Climate change and witch-hunting: the impact of the Little
Ice Age on mentalities. Climate Change, 43, 335–51.
Behringer, W. (2010). A Cultural History of Climate. London: Polity Press.
Bender, M., Sowers, T., & Brook, E. (1997). Gases in ice cores. Proceedings of the
National Academy of Sciences of the United States of America, 94, 8343–9.
Benson, L. V. (2010). Who provided maize to Chaco Canyon after the mid-12th-
century drought?. Journal of Archaeological Science, 37, 621–9.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
0
7

70 Ljungqvist

Benson, L. & Berry, M. S. (2009). Climate change and cultural response in the pre-
historic American Southwest. Kiva, 75, 89–119.
Benson, L., Kashgarian, M., Rye, R., Lund, S., Paillet, F., Smoot, J., Kester, C.,
Mensing, S., Meko, D., & Landström, S. (2002). Holocene multidecadal
and multicentennial droughts affecting Northern California and Nevada.
Quaternary Science Reviews, 21, 659–82.
Benson, L., Petersen, K., & Stein, J. (2007). Anasazi (pre-Columbian Native
American) migrations during the middle-12th and late-13th centuries – Were
they drought induced?. Climatic Change, 83, 187–213.
Berger, A. & Loutre, M. F. (1991). Insolation values for the climate of the last
10 million years. Quaternary Science Reviews, 10, 297–317.
Blinman, E. (2008). 2000 years of cultural adaptation to climate change in the
Southwestern United States. AMBIO: A Journal of the Human Environment,
37, 489–97.
Bocinsky, R. K. & Kohler, T. A. (2014). A 2,000-year reconstruction of the rain-fed
maize agricultural niche in the US Southwest. Nature Communications, 5,
5618. doi: 10.1038/ncomms6618.
Bond, G., Kromer, B. & Beer, J., Muscheler, R., Evans. M. N., Showers, W., Hoffmann,
S., Lotti-Bond, R., Hajdas, I., & Bonani, G. (2001). Persistent solar influence on
North Atlantic climate during the Holocene. Science, 294, 2130–6.
Bond, G., Showers, W. & Cheseby, M., Lotti, W., Almasi, P., Demenocal, P. B.,
Priore, P., Cullen, H., Hajdas, I., & Bonani, G. (1997). A pervasive millennial-
scale cycle in North Atlantic Holocene and glacial climates. Science, 278,
1257–66.
Bond, G. C., Showers, W., Elliot, M., Evans, M., Lotti, R., Hajdas, I., Beosnan, G.,
& Johnson, S. (1999). The North Atlantic’s 1–2 kyr climate rhythm’s relation
to Heinrich Events, Dansgaard/Oeschger Cycles and the Little Ice Age. In P.
U. Clark, R. S. Webb, & L. D. Keigwin, eds., Mechanisms of Global Climate
Change at Millennial Time Scales. Washington, DC: Geophysical Monograph
Series. American Geophysical Union, pp. 35–58.
Bradley, R. S. (1999). Paleoclimatology: Reconstructing Climates of the
Quaternary. San Diego, CA: Academic Press.
Brázdil, R., Pfister, C., Wanner, H., von Storch, H., & Luterbacher, J. (2005).
Historical climatology in Europe – The state of the art. Climatic Change, 70,
363–430.
Breitenmoser, P., Beer, J., & Brönnimann, S. (2012). Solar and volcanic finger-
prints in tree-ring chronologies over the past 2000 years. Palaeogeography,
Palaeoclimatology, Palaeoecology, 313–14, 127–39.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
1
7

Human and Societal Dimensions of Past Climate Change 71

Broecker, W. S. & Putnam, A. E. (2013). Hydrologic impacts of past shifts of Earth’s


thermal equator offer insight into those to be produced by fossil fuel CO2.
Proceedings of the National Academy of Sciences of the United States of
America, 110, 16710–15.
Brooke, J. L. (2014). Climate Change and the Course of Global History: A Rough
Journey. New York: Cambridge University Press.
Bryson, R. A. & Murray, T. J. (1977). Climates of Hunger: Mankind and the World’s
Changing Weather. Madison: University of Wisconsin Press.
Buckland, P. C., Amorosi, T. & Barlow, L. K., Dugmore, A. J., Mayewski,
P. A., McGovern, T. H., Ogilvie, A. E. J., Sadler, J. P., & Skidmore, P. (1996).
Bioarchaeological and climatological evidence for the fate of Norse farmers in
medieval Greenland. Antiquity, 70, 88–96.
Buhaug, H. (2010). Climate not to blame for African civil wars. Proceedings of
the National Academy of Sciences of the United States of America, 107 (38),
16477–82.
Büntgen, U., Myglan, V. & Ljungqvist, F. C. McCormic, M., Di Cosmo, N., Sigl,
M., Jungclaus, J. H., Wagner, S., Krusic, P. J., Esper, J., Kaplan, J. O., De Vaan,
M. A. C., Luterbacher, J., Wacker, L., Tegel, W., & Kirdyanov. (2016). Cooling
and societal change during the Late Antique Little Ice Age from 536 to around
660 AD. Nature Geoscience, 9, 231–6.
Campbell, B. M. S. (2016). The Great Transition: Climate, Disease and Society in
the Late-Medieval World. Cambridge: Cambridge University Press.
Carey, M. (2012). Climate and history: a critical review of historical climatology
and climate change historiography. Wiley Interdisciplinary Reviews: Climate
Change, 3, 233–49.
Chen, J., Chen, F. & Feng, S., Huang, W., Liu, J., & Zhou, A. (2015). Hydroclimatic
changes in China and surroundings during the Medieval Climate Anomaly
and Little Ice Age: spatial patterns and possible mechanisms. Quaternary
Science Reviews, 107, 98–111.
Chen, J., Rao, Z., Liu, J., Huang, W., Feng, S., & Dong, G. (2016). On the timing
of the East Asian summer monsoon maximum during the Holocene – Does
the speleothem oxygen isotope record reflect monsoon rainfall variability?
Science China Earth Sciences, 59, 2328–38.
Chen, Q. (2015). Climate shocks, dynastic cycles and nomadic conquests: evi-
dence from historical China. Oxford Economic Papers, 67, 185–204.
Cheyette, F. L. (2008). The disappearance of the ancient landscape and the climatic
anomaly of the early Middle Ages: a question to be pursued. Early Medieval
Europe, 16, 127–65.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
2
7

72 Ljungqvist

Christiansen, B. & Ljungqvist, F. C. (2012). The extra-tropical Northern


Hemisphere temperature in the last two millennia: reconstructions of low-
frequency variability. Climate of the Past, 8, 765–86.
Christiansen, B. & Ljungqvist, F. C. (2017). Challenges and perspectives for large-
scale temperature reconstructions of the past two millennia. Reviews of
Geophysics, 55, 40–96.
Chu, C. Y. C. & Lee, R. D. (1994). Famine, revolt, and the dynastic cycle–population
dynamics in historical China. Journal of Population Economics, 7, 351–78.
Compo, G. P., Whitaker, J. S., Sardeshmukh, P. D., Matsui, N., Allan, R. J., Yin,
X., Gleason, Jr., B. E., Vose, R. S., Rutledge, G., Bessemoulin, P., Brönnimann,
S., Brunet, M., Crouthamel, R. I., Grant, A. N., Groisman, P. Y., Jones, P. D.,
Kruk, M. C., Kruger, A. C., Marshall, G. J., Maugeri, M., Mok, H. Y., Nordli,
Ø., Ross, T. F., Trigo, R. M., Wang, X. L., Woodruff, S. D., & Worley, S. J.
(2011). The twentieth century reanalysis project. Quarterly Journal of the
Royal Meteorological Society, 137, 1–28.
Cook, B. I., Smerdon, J. E., Seager, R., & Coats, S. (2014). Global warming and 21st
century drying. Climate Dynamics, 43, 2607–27.
Cook, E. R., Anchukaitis, K. J., Buckley, B. M., Jacoby, G. C., & Wright, W. E.
(2010). Asian monsoon failure and megadrought during the last millennium.
Science, 328, 486–9.
Cook, E. R., Briffa, K. R., Meko, D. M., Graybill, D. A., & Funkhouser, G. (1995).
The ‘segment length curse’ in long tree-ring chronology development for pal-
aeoclimatic studies. The Holocene, 5, 229–37.
Cook, E. R., Woodhouse, C. A., & Eakin, C. M. (2004). Long term aridity changes
in the western United States. Science, 306, 1015–18.
Crumley, C. L. (1993). Historic ecotonal shifts. Ecological Applications, 3, 377–84.
Cullen, H., deMenocal, P. B. & Hemming, S., Hemming, G., Brown, F. H.,
Gulderson, T., & Sirocko, F. (2000). Climate change and the collapse of the
Akkadian empire: evidence from the deep sea. Geology, 28, 379–82.
d’Alpoim Guedes, J. A., Crabtree, S. A., Bocinsky, R. K., & Kohler, T. A. (2016).
Twenty-first century approaches to ancient problems: climate and society.
Proceedings of the National Academy of Sciences of the United States of
America, 113, 14483–91.
D’Arrigo, R., Frank, D., Jacoby, G., & Pederson, N. (2001). Spatial response to
major volcanic events in or about AD 536, 934 and 1258: frost rings and other
dendrochronological evidence from Mongolia and northern Siberia. Climatic
Change, 49, 239–46.
deMenocal, P. B. (2001). Cultural responses to climate change during the Late
Holocene. Science, 292, 667–73.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
3
7

Human and Societal Dimensions of Past Climate Change 73

Diamond, J. (2005). Collapse: How Societies Choose to Fail or Succeed.


New York: Penguin Books.
Diaz, H. & Trouet, V. (2014). Some perspectives on societal impacts of past cli-
matic changes. History Compass, 12, 160–77.
Douglas, P. M. J., Demarest, A. A., Brenner, M., & Canuto, M. A. (2016). Impacts of
climate change on the collapse of lowland Maya civilization. Annual Review
of Earth and Planetary Sciences, 44, 613–45.
Dugmore, A. J., Keller, C., & McGovern, T. H. (2007). Norse Greenland settle-
ment: reflections on climate change, trade, and the contrasting fates of human
settlements in the North Atlantic islands. Arctic Anthropology, 44, 12–36.
Dykoski, C. A., Edwards, R. L., Chen, H., Yuan, D., Cai, Y., Zhang, M., Lin, Y.,
An, Z., & Revenaugh, J. (2005). A high-resolution, absolute-dated Holocene
and deglacial Asian monsoon record from Dongge Cave, China. Earth and
Planetary Science Letters, 233, 71–86.
Esper, J., Cook, E. R., Krusic, P. J., Peters, K., & Schweingruber, F. H. (2003). Tests
of the RCS method for preserving low-frequency variability in long tree-ring
chronologies. Tree-Ring Research, 59, 81–98.
Esper, J., Krusic, P. J., Ljungqvist, F. C., Luterbacher, J., Carrer, M., Cook, E., Davi,
N. K., Hartl-Marier, C., Kirdyanov, A., Konter, O., Myglan, V., Timonen, M.,
Treydte, K., Truet, V., Villalba, R., Yang, B., & Büntgen, U. (2016). Ranking
of tree-ring based temperature reconstructions of the past millennium.
Quaternary Science Reviews, 145, 134–51.
Fagan, B. M. (2000). The Little Ice Age: How Climate Made History, 1300–1850.
New York: Basic Books.
Fairchild, I. J. & Baker, A. (2012), Speleothem Science: From Process to Past
Environments. Oxford: Wiley.
Fan, K.-W. (2010). Climatic change and dynastic cycles in Chinese history: a
review essay. Climatic Change, 101, 565–73.
Fan, K.-W. (2015). Climate change and Chinese history: a review of trends, topics,
and methods. Wiley Interdisciplinary Reviews: Climate Change, 6, 225–38.
Fang, J. & Liu, G. (1992). Relationship between climatic change and the nomadic
southward migrations in eastern Asia during historical times. Climatic
Change, 22, 1511–69.
Finné, M., Holmgren, K., Sundqvist, H. S., Weiberg, E., & Lindblom, M.
(2011): Climate in the eastern Mediterranean, and adjacent regions, during the
past 6000 years – A review. Journal of Archaeological Science, 38, 3153–73.
Fritts, H. (1976). Tree Rings and Climate. London: Academic Press.
Fyfe, J. C., Meehl, G. A. & England, M. H., Mann, M. E., Santer, B. D., Flato, G.
M., Hawkins, E., Gillet, N. P., Xie, S.-P., Kosaka, Y., & Swart, N. C. (2016).

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
4
7

74 Ljungqvist

Making sense of the early-2000s warming slowdown. Nature Climate Change,


6, 224–8.
Ge, Q., Hao, Z., Zheng, J., & Shao, X. (2013). Temperature changes over the past
2000 yr in China and comparison with the Northern Hemisphere. Climate of
the Past, 9, 1153–60.
Ge, Q., Zheng, J., Tian, Y., Wu, W., Fang, X., & Wang, W.-C. (2008). Coherence of
climatic reconstruction from historical documents in China by different stud-
ies. International Journal of Climatology, 28, 1007–24.
Gill, R. B. (2000). The Great Maya Droughts: Water, Life, and Death. Albuquerque:
University of New Mexico Press.
Glaser, R. (2008). Klimageschichte Mitteleuropas. 1200 Jahre Wetter, Klima,
Katastrophen. Darmstadt: Wissenschaftliche Buchgesellschaft [in German].
Gleisner, H., Thejll, P., Christiansen, B., & Nielsen, J. K. (2015). Recent global
warming hiatus dominated by low-latitude temperature trends in surface and
troposphere data. Geophysical Research Letters, 42, 510–17. doi: 10.1002/
2014GL062596.
Gray, L. J., Beer, J. & Geller, M., Haigh, J. D., Lockwood, M., Matthes, K.,
Cubasch, U., & Fleitmann, D. (2010). Solar influences on climate. Reviews of
Geophysics, 48, RG4001. doi: 10.1029/2009RG000282.
Gräslund, B. & Price, N. (2012). Twilight of the gods? The ‘dust veil event’ of AD
536 in critical perspective. Antiquity, 86, 428–43.
Gunn, J. D., ed. (2000). The Years Without Summer. Tracing A.D. 536 and its
Aftermath. Oxford: Archaeopress.
Halstead, P. & O’Shea, J., eds. (1989). Bad Year Economics: Cultural Responses to
Risk and Uncertainty. New York: Cambridge University Press.
Hansen, J., Ruedy, R., Sato, M., & Lo, K. (2010). Global surface temperature change.
Reviews of Geophysics, 48, RG4004. doi: 10.1029/2010RG000345.
Hansen, J., Sato, M. & Ruedy, R., Lo, K., Lea, D. W., & Medina-Elizade, M. (2006).
Global temperature change. Proceedings of the National Academy of Sciences
of the United States of America, 103, 14288–93.
Hao, Z., Zheng, J., Ge, Q., & Zhang, X. (2012). Spatial patterns of precipitation
anomalies for 30-yr warm periods in China during the past 2000 years. Acta
Meteorologica Sinica, 26, 278–88.
Hao, Z., Zheng, J. & Zhang, X., Liu, H., Li, M., & Ge, Q. (2016). Spatial patterns
of precipitation anomalies in eastern China during centennial cold and warm
periods of the past 2000 years. International Journal of Climatology, 26,
467–75.
Held, I. M. & Soden, B. J. (2006). Robust responses of the hydrological cycle to
global warming. Journal of Climate, 19, 5686–99.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
5
7

Human and Societal Dimensions of Past Climate Change 75

Hellmann, L., Nikolaev, A. & Ljungqvist, F. C., Churakova (Sidorova), O., Düthorn,
E., Esper, J., Hülsmann, L., Kirdyanov, A. V., Moiseev, P., & Myglan, V. S.
(2016). Diverse growth trends and climate responses across Eurasia’s boreal
forest. Environmental Research Letters, 11, 074021. doi: 10.1088/1748–9326/
11/7/074021.
Hiller, A., Boettger, T., & Kremenetski, C. (2001). Medieval climatic warming
recorded by radiocarbon dated alpine tree-line shift on the Kola Peninsula,
Russia. The Holocene, 11, 491–7.
Hind, A., Zhang, Q., & Brattström, G. (2016). Problems encountered when defin-
ing Arctic amplification as a ratio. Scientific Reports, 6, 30469. doi: 10.1038/
srep30469.
Holmes, J. A. (2008). How the Sahara became dry. Science, 320, 752–3.
Hsiang, S. M. & Burke, M. (2014). Climate, conflict, and social stability: what does
the evidence say? Climate Change, 123, 39–55.
Hsiang, S. M., Burke, M., & Miguel, E. (2013). Quantifying the influence
of climate on human conflict. Science, 341, 1235367. doi: 10.1126/
science.1235367.
Hsu, K. J. (1998). Sun, climate, hunger, and mass migration. Science in China
Series D. Earth Sciences, 41, 449–72.
Hughes, M. K. & Graumlich, L. J. (1996). Multi-millennial dendroclimatic studies
from the western United States. In P. D. Jones, R. S. Bradley, & J. Jouzel, eds.,
Climatic Variations and Forcing Mechanisms of the Last 2000 Years. Volume
141. NATO ASI Series, pp. 109–24.
Huang, S. P., Pollack, H. N., & Shen, P. Y. (2008). A late Quaternary climate recon-
struction based on borehole heat flux data, borehole temperature data, and the
instrumental record. Geophysical Research Letters, 35, L13703. doi: 10.1029/
2008GL034187.
Hugo, G. & Currey, B. (1989). Famine: As a Geographical Phenomenon. Dordrecht,
Springer.
Huntington, E. (1907). The Pulse of Asia: A Journey in Central Asia Illustrating
the Geographic Basis of History. Boston: Houghton, Mifflin and Company.
Huntington, E. (1913). Changes of climate and history. American Historical
Review, 19, 213–32.
IPCC. (2013). Climate Change 2013: The Physical Science Basis. Contribution of
Working Group I to the Fifth Assessment Report of the Intergovernmental
Panel on Climate Change [T. F. Stocker, D. Qin, G. K. Plattner, M. Tignor,
S. K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex, & P. M. Midgley, eds.].
Cambridge, United Kingdom and New York, NY, USA: Cambridge University
Press. doi: 10.1017/CBO9781107415324.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
6
7

76 Ljungqvist

Issar, A. & Zohar, M. (2004). Climate Change: Environment and Civilization in


the Middle East. Berlin: Springer.
Jones, P. (2016). The reliability of global and hemispheric surface temperature
records. Advances in Atmospheric Sciences, 33, 269–82.
Jones, P. D., Briffa, K. R. & Osborn, T. J., Loughm J. M., & van Ommen, T. D.
(2009). High-resolution palaeoclimatology of the last millennium: a review of
current status and future prospects. The Holocene, 19, 3–49.
Jouzel, J. (2013). A brief history of ice core science over the last 50 yr. Climate of
the Past, 9, 2525–47.
Karlén, W. & Kuylenstierna, J. (1996). On solar forcing of Holocene climate: evi-
dence from Scandinavia. The Holocene, 6, 359–36.
Karlén, W. & Larsson, L. (2007). Mid-Holocene climatic and cultural dynamics
in Northern Europe. In D. G. Anderson, K. Maasch, & D. H. Sandweiss, eds.,
Climate Change and Cultural Dynamics: A Global Perspective on Mid-
Holocene Transitions. London: Academic Press, pp. 407–34.
Klomp, J. & Bulte, E. (2013). Climate change, weather shocks, and violent con-
flict: a critical look at the evidence. Agricultural Economics, 44, 63–78.
Krämer, D. (2015). Menschen Grasten nun mit dem Vieh. Die Letzte Grosse
Hungerkrise der Schweiz 1816/17. Basel: Schwabe [in German].
Kullman, L. (2015). Higher-than-present Medieval pine (Pinus sylvestris L.) tree-
line along the Swedish Scandes. Landscape Online, 42, 1–14.
Kuper, R. & Kröpelin, S. (2006). Climate-controlled Holocene occupation in the
Sahara: motor of Africa’s evolution. Science, 313, 803–7.
Laird, K. R., Fritz, S. C., Maasch, K. A., & Cumming, B. F. (1996). Greater drought
intensity and frequency before AD 1200 in the northern Great Plains, USA.
Nature, 384, 552–4.
Lamb, H. H. (1972–7). Climate: Present, Past and Future 1–2. London: Methuen.
Larsen, L. B., Vinther, B. M. & Briffa, K. R., Melvin, T. M., Clausen, H. B., Jones, P.
D., Andersen, M. L. S., Hammer, C. U., Eronen, M., Grudd, H., Gunnarson, B.
E., Hantemirov, R. M., Naurzbaev, M. M., & Nicolussi, K. (2008). New ice core
evidence for a volcanic cause of the A.D. 536 dust veil. Geophysical Research
Letters, 35, L04708. doi: 10.1029/2007GL032450.
Leduc, G., Schneider, R., Kim, J.-H., & Lohmann, G. (2010). Holocene and Eemian
sea surface temperature trends as revealed by Alkenone and Mg/Ca paleother-
mometry. Quaternary Science Reviews, 29, 989–1004.
Lee, H. F. (2014). Climate-induced agricultural shrinkage and overpopulation in
late imperial China. Climate Research, 59, 229–42.
Lee, H. F. & Zhang, D. D. (2015). Quantitative analysis of climate change and
human crises in history. In M.-P. Kwan, D. Richardson, D. Wang, & C. Zhou,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
7

Human and Societal Dimensions of Past Climate Change 77

eds., Space-Time Integration in Geography and GIScience: Research Frontiers


in the US and China. Dordrecht: Springer, pp. 235–67.
Liu, Y., Cai, Q. F., Song, H. M., An, Z. S., & Linderholm, H. W. (2011). Amplitudes,
rates, periodicities and causes of temperature variations in the past 2485 years
and future trends over the central-eastern Tibetan Plateau. Chinese Science
Bulletin, 6, 2986–94.
Ljungqvist, F. C. (2009). Global nedkylning: Klimatet och människan under
10 000 år. Stockholm: Norstedts [in Swedish].
Ljungqvist, F. C. (2011). The spatio-temporal pattern of the Mid-Holocene Thermal
Maximum. Geografie, 116, 91–110.
Ljungqvist, F. C., Krusic, P. J., Brattström, G., & Sundqvist, H. S. (2012). Northern
Hemisphere temperature patterns in the last 12 centuries. Climate of the
Past, 8, 227–49.
Ljungqvist, F. C., Krusic, P. J. & Sundqvist, H. S., Zorita, E., Brattström, G., &
Frank, D. (2016). Northern Hemisphere hydroclimatic variability over the past
twelve centuries. Nature, 532, 94–8.
Löwenborg, D. (2012). An Iron Age shock doctrine: did the AD 536–7 event trigger
large-scale social changes in the Mälaren valley area? Journal of Archaeology
and Ancient History, 4, 1–29.
Lucero, L. J. (2006). Water and Ritual: The Rise and Fall of Classic Maya Rulers.
Austin: University of Texas Press.
Luterbacher, J. & Pfister, C. (2015). The year without a summer. Nature
Geoscience, 8, 246–8.
Luterbacher, J., Werner, J. P. & Smerdon, J., Fernández-Donado, L., González-
Rouco, F. J., Barriopedro, D., Ljungqvist, F. C, Büntgen, U., Zorita, E., Wagner,
S., Esper, J., McCarroll, D., Toreti, A., Frank, D., Jungclaus, J. H., Barriendos,
M., Bertolin, C., Bothe, O., Brázdil, R., Camuffo, D., Dobrovolný, P., Gagen,
M., García-Bustamante, E., Ge, Q., Gómez-Navarro, J. J., Guiot, J., Hao, Z.,
Hegerl, G. C., Holmgren, K., Klimenko, V. V., Martín-Chivelet, J., Pfister, C.,
Roberts, N., Schindler, A., Schurer, A., Solomina, O., von Gunten, L., Wahl,
E., Wanner, H., Wetter, O., Xoplaki, E., Yuan, N., Zanchettin, D., Zhang, H.,
& Zerefos, C. (2016). European summer temperatures since Roman times.
Environmental Research Letters, 11, 024001. doi: 10.1088/1748–9326/11/1/
024001.
Madsen, C. K. (2014) ‘Pastoral Settlement, Farming, and Hierarchy in Norse
Vatnahverfi, South Greenland’. Copenhagen: University of Copenhagen,
unpublished PhD thesis.
Mann, M. E., Zhang, Z. & Hughes, M. K., Bradley, R. S., Miller, S. K., Rutherford,
S., & Ni, F. (2008). Proxy-based reconstructions of hemispheric and global

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
8
7

78 Ljungqvist

surface temperature variations over the past two millennia. Proceedings of the
National Academy of Sciences of the United States of America, 105, 13252–7.
Mann, M. E., Zhang, Z. & Rutherford, S., Bradley, R. S., Hughes, M. K., Shindell,
D., Ammann, C., Faluvegi, G., & Ni, F. (2009). Global signatures and dynami-
cal origins of the Little Ice Age and Medieval Climate Anomaly. Science, 326,
1256–60.
Matthews, J. A. & Briffa, K. R. (2005). The ‘Little Ice Age’: re-evaluation of an
evolving concept. Geografiska Annaler, 87A, 17–36.
Mayewski, P. A., Rohling, E. E. & Stager, J. C., Karlén, W., Maasch, K. A., Meeker,
L. D., Meyerson, E. A., Gasse, F., van Krevels, S., Holmgren, K., Lee-Thorp, J.,
Rosqvist, G., Rack, F., Staubvasser, M., Schneider, R. R., & Steig, E. J. (2004).
Holocene climate variability. Quaternary Research, 62, 243–55.
Mazepa, V. S. (2005). Stand density in the last millennium at the upper tree-line
ecotone in the Polar Ural Mountains. Canadian Journal of Forest Research,
35, 2082–91.
McCormick, M., Büntgen, U. & Cane, M. A., Cook, E. R., Harper, K., Huyers, P.,
Litt, T., Manning, S. W., Mayewski, P. A., More, A. F. M., Nicolussi, K., & Tegel,
W. (2012). Climate change during and after the Roman Empire: reconstructing
the past from scientific and historical evidence. Journal of Interdisciplinary
History, 4, 169–220.
McMichael, A. J. (2012). Insights from past millennia into climatic impacts on
human health and survival. Proceedings of the National Academy of Sciences,
109, 4730–7.
McNeill, J. R. (2016). Historians, superhistory, and climate change. In A. Jarrick,
J. Myrdal, & M. Wallenberg Bondesson, eds., Methods in World History:
A Critical Approach. Lund: Nordic Academic Press, pp. 19–43.
Meierding, E. (2013) Climate change and conflict: avoiding small talk about the
weather. International Studies Review, 15, 185–203.
Melander, E., Pettersson, T., & Themnér, L. (2016). Organized violence, 1989–
2015. Journal of Peace Research, 53, 727–42.
Melvin, T. M. & Briffa, K. R. (2008). A ‘signal-free’ approach to dendroclimatic
standardisation. Dendrochronologia, 26, 71–86.
Moberg, A., Sonechkin, D. M., Holmgren, K., Datsenko, N. M., & Karlén, W.
(2005). Highly variable Northern Hemisphere temperatures reconstructed
from low- and high-resolution proxy data. Nature, 433, 613–17.
Morice, C. P., Kennedy, J. J., Rayner, N. A., & Jones, P. D. (2012). Quantifying
uncertainties in global and regional temperature change using an ensemble
of observational estimates: The HadCRUT4 data set. Journal of Geophysical
Research: Atmospheres, 117, D08101. doi: 10.1029/2011JD017187.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
9
7

Human and Societal Dimensions of Past Climate Change 79

Nelson, M. C., Ingram, S. E. & Dugmore, A. J., Streeter, R., Peeples, M. A.,
McGowern, T. H., Hegmin, M., Arneborg, J., Kintigh, K., Brewington,
S., Spielmann, K. A., Simpson, I. A., Strawhacker, C., Comeau, L. E. L.,
Torvinen, A., Madsen, C. K., Hambrecht, G., & Smiarowski, K. (2016).
Climate challenges, vulnerabilities, and food security. Proceedings of
the National Academy of Sciences of the United States of America, 113,
298–303.
Newman, L. F., ed. (1990). Hunger in History: Food Shortage, Poverty, and
Deprivation. Cambridge/Oxford: Blackwell.
O’Loughlin, J., Linke, A. M., & Witmer, F. D. (2014). Modeling and data choices
sway conclusions about climate–conflict links. Proceedings of the National
Academy of Sciences of the United States of America, 111, 2054–5.
PAGES 2k Consortium (2013). Continental-scale temperature variability during
the past two millennia. Nature Geoscience, 6, 339–46.
Parker, G. (2013). Global Crisis: War, Climate and Catastrophe in the Seventeenth
Century. New Haven, CT: Yale University Press.
Parry, M. L. (1978). Climatic Change, Agriculture and Settlement. Folkestone:
William Dawson & Sons.
Parry, M. L. & Carter, T. R. (1983). Assessing Impacts of Climatic Change in Marginal
Areas: The Search for an Appropriate Methodology. Laxenburg: International
Institute for Applied Systems Analysis.
Parry, M. L., Carter, T. R., & Konijn, N. T., eds. (1988a). The Impact of Climatic
Variations on Agriculture. Volume 1. Assessments in Cool Temperate and
Cold Regions. Dordrecht: Kluwer.
Parry, M. L., Carter, T. R., & Konijn, N. T., eds. (1988b). The Impact of Climatic
Variations on Agriculture. Volume 2. Assessments in Semi-arid Regions.
Dordrecht: Kluwer.
Payette, S., Filion, L., Delwaide, A., & Begin, C. (1989). Reconstruction of tree-line
vegetation response to long-term climate change. Nature, 341, 429–32.
Pettersson, T. & Wallensteen, P. (2015). Armed conflicts, 1946–2014. Journal of
Peace Research, 52, 536–50.
Pfister, C. (2007). Climatic extremes, recurrent crises and witch hunts: strategies
of European societies in coping with exogenous shocks in the late sixteenth
and early seventeenth centuries. Medieval History Journal, 10, 33–73.
Pierrehumbert, R. T. (2010). Principles of Planetary Climate. Cambridge: Cambridge
University Press.
Post, J. (1985). Food Shortage, Climatic Variability, and Epidemic Disease in
Preindustrial Europe: The Mortality Peak in the Early 1740s. Ithaca, NY:
Cornell University Press.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
0
8

80 Ljungqvist

Raleigh, C., Linke, A., & O’Loughlin, J. (2014). Extreme temperatures and
violence. Nature Climate Change, 4, 76–7.
Raspopov, O. M., Dergachev, V. A. & Esper, J., Kozyreva, O., Frank, D., Ogurtsov,
M. G., Kolström, T., & Shao, X. (2008). The influence of the de Vries (~200-year)
solar cycle on climate variations: results from the central Asian mountains
and their global link. Palaeogeography, Palaeoclimatology, Palaeoecology,
259, 6–16.
Raspopov, O. M., Dergachev, V. A., & Kolström, T. (2004). Periodicity of climate
conditions and solar variability derived from dendrochronological and other
palaeo-climatic data in high latitudes. Palaeogeography, Palaeoclimatology,
Palaeoecology, 209, 127–39.
Renssen, H., Seppä, H., Heiri, O., Roche, D. M., Goosse, H., & Fichefet, T. (2009).
The spatial and temporal complexity of Holocene thermal maximum. Nature
Geoscience, 2, 411–14.
Robock, A. (2000). Volcanic eruptions and climate. Reviews of Geophysics, 38,
191–219.
Rohde, R., Muller, R. A. & Jacobsen, R., Muller, R., Perlmutter, S., Rosefelt, A.
Wurtele, J., Groom, D., & Wickham, C. (2013). A new estimate of the average
earth surface land temperature spanning 1753 to 2011. Geoinformatics and
Geostatistics: An Overview, 1, 1. doi: 10.4172/gigs.1000101.
Rosen, A. M. (2007). Civilizing Climate: Social Responses to Climate Change in
the Ancient Near East. Lanham, MD: Altamira Press.
Rotberg, R. & Rabb, T., eds. (1985). Hunger in History: The Impacts of Changing Food
Production and Consumption Patterns on Society. Cambridge: Cambridge
University Press.
Sandweiss, D. H., Maasch K. A., & Anderson, D. G. (1999). Transitions in the mid-
Holocene. Science, 283, 499–500.
Scheffer, M., Carpenter, S., Foley, J. A., Folke, C., & Walker, B. (2001). Catastrophic
shifts in ecosystems. Nature, 413, 591–6.
Scheffran, J., Brzoska, M., Brauch, H. G., Link, P. M., & Schilling, J. eds. (2012).
Climate Change, Human Security and Violent Conflict: Challenges for
Societal Stability. Berlin: Springer.
Schneider, T., Bischoff, T., & Haug, G. H. (2014). Migrations and dynamics of the
intertropical convergence zone. Nature, 513, 45–53.
Schönwiese, C. D. (1995). Klimaänderungen – Daten, Analysen, Prognosen.
Berlin: Springer Verlag [in German].
Sigl, M., Winstrup, M. & McConnell, J. R., Welten, K. C., Plunkett, G., Ludlow, F.,
Büntgen, U., Caffee, M., Chellman, N., Dahl-Jensen, D., Fischer, H., Kipfstuhl,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
1
8

Human and Societal Dimensions of Past Climate Change 81

S., Kostick, C., Maselli, O. J., Mekhaldi, F., Mulvaney, R., Muscheler, R.,
Pasteris, D. R., Pilcher, J. R., Salzer, M., Schüpbach, S., Steffensen, J. P.,
Vinther, B. M., & Woodruff, T. E. (2015). Timing and climate forcing of vol-
canic eruptions for the past 2,500 years. Nature, 523, 543–9.
Sinha, A., Berkelhammer, M., Stott, L., Mudelsee, M., Cheng, H., & Biswas, J.
(2011). The leading mode of Indian Summer Monsoon precipitation variabil-
ity during the last millennium. Geophysical Research Letters, 38, L15703.
doi: 10.1029/2011GL047713.
Slavin, P. (2016). Climate and famines: a historical reassessment. WIREs Climate
Change, 7, 433–47.
Solomina, O. N., Bradley, R. S. & Jomelli, V., Geirsdottir, A., Kaufman, D. S.,
Koch, J., McKay, N. P., Masiokas, M., Miller, G., Nesje, A., Nicolussi, K.,
Owen, L. A., Putnam, A. E., Wanner, H., Wiles, G., & Yang, B. (2016). Glacier
fluctuations during the past 2000 years. Quaternary Science Reviews, 149,
61–90.
Sorokin, P. (1975). Hunger as a Factor in Human Affairs Gainesville: University
Presses of Florida.
St. George, S. (2014). An overview of tree-ring width records across the Northern
Hemisphere. Quaternary Science Reviews, 95, 132–50.
Steinhilber, F., Abreua, J. A. & Beer, J., Brunner, I., Christi, M., Fisher, H., Heikkilä,
U., Kubik, O. W., Mann, M., McCracken, K., Miller, H., Miyahara, H., Oerter,
H., & Wilhelms, F. (2012). 9,400 years of cosmic radiation and solar activ-
ity from ice cores and tree rings. Proceedings of the National Academy of
Sciences of the United States of America, 109, 5967–71.
Stothers, R. B. (1999). Volcanic dry fogs, climate cooling, and plague pandemics in
Europe and the Middle East. Climatic Change, 42, 713–23.
Su, Y., Liu, L., Fang, X. Q., & Ma, Y. N. (2016). The relationship between cli-
mate change and wars waged between nomadic and farming groups from the
Western Han Dynasty to the Tang Dynasty period. Climate of the Past, 12,
137–50.
Taricco, C., Mancuso, S., Ljungqvist, F. C., Alessio, S., & Ghil, M. (2015).
Multispectral analysis of Northern Hemisphere temperature records over the
last five millennia. Climate Dynamics, 45, 83–104.
Tierney, J. E., Pausata, F. S. R., & deMenocal, P. B. (2017). Rainfall regimes of the
Green Sahara. Science Advances, 3, e1601503. doi: 10.1126/sciadv.1601503.
Toohey, M., Krüger, K. & Sigl, M., Stordal, F., & Svensen, H. (2016). Climatic and
societal impacts of a volcanic double event at the dawn of the Middle Ages.
Climatic Change, 136, 401–12.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
2
8

82 Ljungqvist

Tvauri, A. (2014). The impact of the climate catastrophe of 536–537 AD in Estonia


and neighbouring areas. Estonian Journal of Archaeology, 18, 30–56.
Wagner, G., Beer, J. & Masarik, J., Muscheler, R., Kubik, P. W., Mende, W.,
Laj, C., Raisbeck, G. M., & Yiou, F. (2001). Presence of the solar de Vries
cycle (~205 years) during the last ice age. Geophysical Research Letters,
28, 303–6.
Walter, J. & Schofield, R., eds. (1989). Famine, Disease and the Social Order in
Early Modern Society. Cambridge: Cambridge University Press.
Wanner, H., Beer, J. & Bütikofer, J., Crowley, T. J., Cubasch, U., Flückinger, J.,
Goosse, H., Grosjean, M., Joos, F., Kaplan, J. O., Küttel, M., Müller, S.
A., Prentice, I. C., Solomina, O., Stocker, T. F., Tarasov, P., Wagner, M. &
Widmann, M. (2008). Mid- to Late Holocene climate change: an overview.
Quaternary Science Reviews, 27, 1791–1828.
Wanner, H., Solomina, O., Grosjean, M., Ritz, S., & Jetel, M. (2011). Structure and
origin of Holocene cold events. Quaternary Science Reviews, 30, 3109–23.
Widgren, M. (2012). Climate and causation in the Swedish Iron Age: learning from
the present to understand the past. Geografisk Tidskrift, 112, 126–34.
Willard, D. A., Bernhardt, C. E., Korejwo, D. A., & Meyers, S. R. (2005). Impact
of millennial-scale Holocene climate variability on eastern North American
terrestrial ecosystems: pollen-based climatic reconstruction. Global and
Planetary Change, 47, 17–35.
Wilson, R., Anchukaitis, K. & Briffa, K. R., Büntgen, U., Cook, E., D’Arrigo, R.,
Davi, N., Esper, J., Frank, D., Gunnarsson, B., Hegerl, G., Helama, S., Klesse,
S., Krusic, P. J., Linderholm, H. W., Myglan, V., Osborn, T. J., Rydval, M.,
Schneider, L., Schurer, A., Wiles, G., Zhang, P., & Zorita, E. (2016). Last mil-
lennium northern hemisphere summer temperatures from tree rings: Part
I: The long term context. Quaternary Science Reviews, 134, 1–18.
Yang, B., Kang, S., Ljungqvist, F. C., Zhao, Y., He, M., & Qin, C. (2014). Drought
variability at the northern fringe of the Asian summer monsoon region over
the past millennia. Climate Dynamics, 43, 845–59.
Zhang, D., Lee, H. F., Wang, C., Lie, B., Pei, Q., Zhang, J., & An, Y. (2011). The
causality analysis of climate change and large-scale human crisis. Proceedings
of the National Academy of Sciences of the United States of America, 108,
17296–301.
Zhang, D. D., Pei, Q. & Lee, H. F., Zhang, J., Chan, C. Q., Li, B., Li, J., & Zhang,
X. (2015). The pulse of imperial China: a quantitative analysis of long-term
geopolitical and climatic cycles. Global Ecology and Biogeography, 24,
187–96.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
3
8

Human and Societal Dimensions of Past Climate Change 83

Zhang, D. D., Zhang, J., Lee, H. F., & He, Y. (2007). Climate change and war
frequency in eastern China over the last millennium. Human Ecology, 35,
403–14.
Zhang, P., Cheng, H. & Edwards, E., Chen, F., Wang, Y., Yang, X., Liu, J., Tan, M.,
Wang, X., Liu, J., An, C., Dai, Z., Zhou, J., Zhang, D., Jia, J., Jin, L., & Johnson,
K. R. (2008). A test of climate, sun, and culture. Science, 322, 940–2.
Zhang, Z., Tian, H. & Cazelles, B., Kausrud, K. L., Bräuning, A., Guo, F., & Stenseth,
C. (2010). Periodic climate cooling enhanced natural disasters and wars in
China during AD 10–1900. Proceedings of the Royal Society B: Biological
Sciences, 277, 3745–53.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.003
4
8

4 Rural Communities and


Traditional Ecological
Knowledge
Anamaria Iuga, Anna Westin, Bogdan Iancu,
Monica Stroe, and Håkan Tunón

4.1 Introduction
Rural communities, where the population is dependent on agriculture
and animal husbandry, have developed in close relationship with the
local environment. Such interaction may be expressed in local and
traditional knowledge, beliefs, land-use practices, and other forms of
intangible heritage, but also tangibly in the landscape as imprints in
ecosystems. Local knowledge is a vital part of culture, actively used
in the adaptive management of natural resources, and although some
parts may have lost their usefulness, they may be remembered in the
community.
In this chapter we explore the importance of such traditional
ecological knowledge (TEK) (Berkes et al. 2000) in rural agricultural
communities. Our focus is placed on the emergence, transformation,
and disappearance of this knowledge in relation to ecological and
socio-cultural conditions. Examples of current and past TEK from
Romania and Sweden are used to illustrate: (1) how today’s Romanian
TEK can provide insight for historical ecology in other contexts,
such as Sweden; (2) the implications of change and loss of TEK; and
(3) current re-evaluation of TEK, by farmers and by society as a whole.
The presented results are based on the integration of ethno-
logical, anthropological, ecological, and historical fieldwork and
experiences from Sweden and Romania (mainly the mountain-
ous Maramureș County). Chapter 7 in this volume explores links
between rural communities and their environment using the con-
cept of biocultural diversity, while the practical and ethical aspects
84

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
5
8

Rural Communities & Traditional Ecological Knowledge 85

of working together with people in rural communities are discussed


in Chapter 8.

4.1.1 The Evolution and Dynamics of TEK in Rural


Communities
Although today’s agricultural produce is dispersed via the global
market, the actual production still takes place in local farming com-
munities. A local community is defined by the World Intellectual
Property Organization as ‘the human population in a distinct ecologi-
cal area who depend directly on its biodiversity and ecosystem goods
and services for all or part of their livelihood and who have devel-
oped or acquired traditional knowledge as a result of this dependence,
including farmers, fisherfolk, pastoralists, forest dwellers and others’
(www.wipo.int/tk/en/resources/glossary.html).
Agriculture at a specific time and place is the result of inter-
action between various social and ecological factors. Local farmers
have to deal with and adapt to the complexities of both nature and
society. The ecosystems used are situated within villages, which are
also the closest societal context of daily life. These local communi-
ties, however, interact – culturally, economically, and politically –
with the outside world. Agriculture can therefore be labelled a
‘social-ecological system’ (Berkes & Folke 1998), the resilience of
which is affected by the adaptive capacity of the local community.
This is, in turn, influenced by individuals, institutional regulations,
and the impact of global change on individual lives (Ruiz-Mallén &
Corbera 2013). Therefore, in order to understand local rural com-
munities, individual values, the role of local stakeholders, and the
interaction with politics in the decision-making process for the use
of natural resources must be taken into consideration (Agrawal &
Gibson 1999).
The development and sharing of knowledge is essential for the
subsistence of rural communities. Based on past experience, com-
munity members know how and which elements of nature can be
used, e.g. how to harvest hay for winter fodder (Figures 4.1 and 4.2).

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
6
8

86 Iuga et al.

Figure 4.1 The building of a haystack in Botiza (Romania), 2014.


Photo: Bogdan Iancu.

The result is TEK, which denotes a ‘cumulative body of knowledge,


practice and beliefs, evolving by adaptive processes and handed down
through generations by cultural transmission, about the relation of
living beings (including humans) with one another and the environ-
ment’ (Berkes et al. 2000: 1252). Ecological knowledge helps rural
communities manage the local environment and adapt to changing
conditions. Altered conditions require new ideas and may also make
traditional knowledge obsolete (Oteros-Rozas et al. 2013). Loss of
TEK is a natural component of the dynamics of tradition (Handler &
Linnekin 1984; Glassie 1995), but since most such knowledge only
exists in land-use practices and in the memories of people, loss of
local TEK often also means it is lost for the world. The fear of los-
ing the knowledge and culture of rural farming communities has
repeatedly created a need for ethnological and folkloristic documen-
tation, usually initiated from outside the communities (e.g. Mesnil

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
7
8

Rural Communities & Traditional Ecological Knowledge 87

Figure 4.2 Sharpening the scythe during mowing. Botiza (Romania),


2010.
Photo: Anamaria Iuga.

1997; Österling 2010; Tunón 2015). Such ex situ documentation can


conserve obsolete knowledge, but will always be less detailed than
in situ knowledge, and lacks the important element of capturing
contexts and adaptive processes. Some of the obsolete TEK may be
retained locally if it is reflected in beliefs, oral traditions, and rituals
(Berkes et al. 2000: 1254) and may be relevant for the identity of the
community.
The traditional knowledge of a community can also put social
pressure on individuals to conform to specific manners of work-
ing the land and to use the techniques of the ancestors, since such
knowledge is perceived as ‘true’, having stood the test of time, along-
side other rules and norms. This may create a ‘path dependency’ on

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
8

88 Iuga et al.

practices, counteracting innovation and development (Magnusson &


Ottosson 2009). In other cases, change may, conversely, be facilitated,
e.g. if the societal norm is to incorporate new ideas.
The adaptive processes of knowledge acquisition within com-
munities have, from time to time, been disrupted by societal and
political change imposed from the outside. War, land reform, indus-
trialisation, political regime changes, and new policies have all influ-
enced rural socio-economic and agricultural conditions as well as
the conditions for maintaining TEK in communities. Such factors
can force people to abandon traditions, but can also be incentives
for developing new TEK and uses of land, as well as revitalising old
knowledge and traditions. The following section contains examples
of such processes in both Romania and Sweden.

4.1.2 Local Rural Communities and TEK in a


Changing World
Since the late nineteenth century, Romania has experienced five land
reforms, changing the conditions of ownership and use of land. The
penultimate reform took place during the communist period follow-
ing the Second World War (1949–62), imposing collectivisation and
expropriating agricultural land in order to create collectives or state
farms; the latest reform was de-collectivisation, initiated after the
fall of the communist regime in 1989 (Verdery 2003).
Before collectivisation, local communities were influenced by
relationships to the immediate family, but also connections created
through religious confession and linguistic specificity, as well as agri-
cultural and cultural practice (Kligman & Verdery 2011). Until the
mid-twentieth century, resources and tasks were communal mat-
ters, especially in hilly and mountainous regions (Stahl 1946); this
traditional organisation was profoundly different from the collectivi-
sation of the communist regime. During this process, private lands
belonging to small-scale family farms were merged to larger units
administered under strong state influence, and although more than
2,800 Romanian mountain villages were never collectivised – due to

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
9
8

Rural Communities & Traditional Ecological Knowledge 89

the rough terrain and dispersed settlement pattern – they still had
to contribute part of their produce to the state. Collectivisation was
a painful process and local resistance was often forcibly supressed.
This process triggered dramatic changes within rural communities,
as reflected in village economies, social composition, and local cul-
ture. Farmers who lost their lands were forced to work for collective
farms or in urban industries. As a consequence, traditional means
of organising the social and cultural life of the villages disappeared.
This meant that some rituals associated with agricultural work
were abandoned, either because of mechanisation or the communist
authorities’ resentment of rituals and customs, as they were directly
connected to religion and beliefs and therefore deemed obsolete for
the ‘new man’ (Mihăilescu et al. 1993: 24–9). Collectivisation broke
interpersonal bonds based on the old landownership model and
imposed other relationships, largely dependent on the new property
system (Torsello 2003: 79).
During the communist era, many urban workers still regarded
their old villages as their symbolic spaces of belonging (Szelenyi
1998; Kideckel 2001). The village was a meaningful space, valued
because of family ties and the constant circulation of people between
villages and towns, forming ‘diffuse households’ (Mihăilescu 2000).
This contributed to the survival of many traditional communal
attributes until the recent past, despite forty-nine years of collectivi-
sation (Mihăilescu et al. 1993).
The profound changes involving landownership, community
ties, and identity, as well as the rights to products and opportunities
to influence land use and practices, are all likely causes for change
and loss of TEK, although these have not yet been explicitly stud-
ied. After 1990, the de-collectivisation process returned land to the
villagers. Agriculture has to some extent evolved in new directions,
adapting to new markets and EU membership (after 2007).
Interestingly, de-collectivisation has led to a resumption of
certain pre-communist agricultural practices, especially in terms
of arable land management. This is exemplified in the study of the

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
0
9

90 Iuga et al.

Transylvanian village Aurel Vlaicu, which shows that families who


returned to their farms were ill-equipped to run them efficiently.
Many of these new landowners had been living in urban areas during
the communist era and therefore lost their skills and local networks.
One major obstacle was finding money to invest in new technol-
ogy, not least because of extreme inflation and high interest rates.
Farmers therefore had to resort to old tools and practices, which
many preferred since the intensive communist cultivation with
chemical fertilisers and pesticides was considered detrimental to the
land (Verdery 2003).
Swedish agriculture and rural communities have been subject
to gradual change, and at times more radical technological and soci-
etal changes (Myrdal 1999). Three major land reforms have taken
place since the late eighteenth century, all with the purpose of redis-
tributing village land and making land use more efficient. The last
reform, known as laga skifte, took place in 1827 and entailed the
most radical changes.
Prior to these land reforms, villages had well-regulated com-
mon organisations for the use of private land, as well as commonly
held land and property. In many respects, the laga skifte dissolved the
villages. Farms were moved away from the village centres, most land
became individually owned, and farmers started making their own
decisions without involving the village councils. This land-use reform
became one of the most important factors for agrarian development
in Sweden during the second half of the nineteenth century, involving
livestock breeding, new crop rotation systems that included the cul-
tivation of fodder, artificial fertilizers, and the transformation of hay
meadows and pastures into arable land and forest. As a result, semi-
natural hay meadows and pastures gradually disappeared (Myrdal &
Morell 2011), and TEK related to the use of low-input agriculture and
semi-natural grasslands was slowly lost and replaced by knowledge
linked to the new agricultural practices. A recent interview-based
study in Sweden shows that farmers, to some extent, still consider
the informal education their children receive through traditional

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
1
9

Rural Communities & Traditional Ecological Knowledge 91

knowledge transfer a prerequisite for success, even for modern agri-


cultural production (Einarsson 2015). Adaptive land management is
still essential, although there is less ecological complexity than in
modern Romania or Sweden of the past.
Across Europe, agriculture has been modernised through pro-
cesses such as globalisation, mechanisation, intensification, plant
breeding, and societal change, gradually reducing the number of
people employed. This process has gone much further in Western
Europe than in the eastern parts of the continent. In Romania, due
to the de-collectivisation of the 1990s, a reverse trend has been
observed. Agriculture has not changed due to the loss of TEK, but
TEK has been lost as agriculture has changed. As a result, Swedish
farmers tend to have deep insight into high-intensity farming, while
Romanian farmers have a much richer TEK related to low-intensity
farming and use of semi-natural grasslands. This type of TEK has
presumably existed in Sweden, but has now been lost. The loss
of TEK is usually not a problem for Swedish farmers, unless they
encounter situations when certain knowledge is needed again, e.g. in
the management of ‘traditional’ habitats for conservation purposes.
This will be discussed in more detail later in this chapter (see ‘TEK
devalued and revalued’).

4.2 Agriculture and TEK in Rural Communities


Agriculture in Romania is still, as it was in Sweden until the late
nineteenth century, part of the mixed farming system that domi-
nated a large area of Northern Europe. Mixed farming is based on
permanent cultivation fertilised by manure, winter-stabling of
livestock, harvesting of winter fodder, and summer grazing (Grigg
1974). A key feature is the combination of animal husbandry and
cultivation, often complemented with fishing, hunting, forest pro-
cessing, or other sources of income. Agriculture has some charac-
teristics that can be used as focal points for the understanding of
TEK, as will be illustrated by the Romanian and Swedish examples
that follow.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
2
9

92 Iuga et al.

• Work carried out by single farmers and farming households is based on


decisions made in a societal context, e.g. as some land is communally
owned and as the mentality of the community may influence what
practices are acceptable.
• Resource extraction from the ecosystem depends on a considerable input
of labour, often concentrated in time, providing people with compelling
incentives to improve methods to facilitate work or increase yields.
• Agriculture and agricultural practices are determined by the annual
seasons.
• Agriculture is dependent on the weather. Harvests can be destroyed
by bad weather and the timing of work needs to be adapted to weather
conditions. There is therefore a great need for communities to understand
and be able to predict the weather.

4.2.1 Common Natural Resources


The management and use of pastures is still important in Romanian
mixed agriculture. Many of the pastures are communal and the live-
stock is kept either close to the village centre or at more remote
shielings (stâna), where the animals spend all summer. This trans-
humance system is still in use across Romania. Usually, the pasture
is owned by the community and a head shepherd rents the land and
tends animals belonging to several families. He employs other men
for herding, milking, and the production of cheese.
The use of common pastures is associated with an extensive
TEK related to grazing, herding, and milk processing, but also to a
variety of traditions. In some villages of Maramureș, the sheep-own-
ing families organise ceremonies including ‘milk-measuring feasts’
at the beginning of each grazing season. The purpose is to deter-
mine the family proportion of ‘input’ and ‘output’. In many villages,
a priest conducts a religious ceremony, which is followed by other
symbolic actions aimed to protect the animals and to ensure a good
season, and afterwards the farmers organise picnics with local food
and drink (Figure 4.3). During the grazing season, each family pays for
its share with food and work, with an amount proportional to the size
of its herd, and after the summer, it pays money to the shepherds for

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
3
9

Rural Communities & Traditional Ecological Knowledge 93

Figure 4.3 During the milk-measurement custom in Șurdești village


(Romania), the milk from all animals is measured in order to divide
input and output during the season, 2009.
Photo: Anamaria Iuga.

every animal brought to the stâna. The output includes a variety of


milk products, above all cheese, which are distributed proportionally
among the families, calculated on the basis of the amount of milk
determined at the feast.
Swedish pastoralism also had many communal aspects until
the mid-nineteenth century. Pastures were village, parish, or hun-
dred commons, with clear regulations on how, when, and for what
purposes they could be used, e.g. regarding the length of the grazing
season and the number of livestock allowed (Myrdal 2012: 155–69).
Respecting agreed grazing periods was an important way of prevent-
ing damage to common crops. The northern Swedish shieling (fäbod)
system, which emerged in the sixteenth and seventeenth centur-
ies and peaked during the nineteenth century, was regulated by an

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
4
9

94 Iuga et al.

Figure 4.4 Collective work of mowing. Fundata (Romania), 2014.


Photo: Bogdan Iancu.

elaborate system of rules and institutions (see further in Chapter 7


and Larsson 2009).

4.2.2 Work Exchange


In addition to cooperation on common land, societal interaction in
traditional rural communities also included helping one another
with the more demanding work on private land, especially during
the hay harvest (Figure 4.4). Generally, this kind of help is offered
within smaller groups, such as families and neighbours. One exam-
ple, recorded in Romania, is the practice called claca, where several
members of a community helped one another with mowing, harvest-
ing cereal, or any other work that needed additional hands. Apart
from speeding up the work pace, this practice also maintained social
relations within the community, as it was usually followed by feasts
or celebrations.
Similar work exchanges were common also in Sweden, dur-
ing both harvest and mowing, and included the lending of horses,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
5
9

Rural Communities & Traditional Ecological Knowledge 95

Figure 4.5 The family of Lars Olsson with extra hands is mowing
the meadows of the Småland farm, county of Jämtland, Sweden, 31
July 1916.
Photo: Nils Svensson/ Photo collections of Jamtli Museum.

machinery, or even farm hands (Figure 4.5). Such exchanges were


built on reciprocity (Westin et al. 2017). As in Romania, this type
of collaboration was often accompanied by celebrations or feasts,
e.g. the ‘mowing-porridge’ (slåttergröt), or ‘mowing-beer’ (slåtteröl)
(e.g. Persson et al. 1996). It is likely that such work exchanges, in
Romania as well as in Sweden, have been useful for sharing the work-
load as well as creating tighter bonds within the local communities.

4.2.3 Seasonality and Timing


A lot of TEK concerns the seasonal cycle of agriculture, the impor-
tance of which can be seen in the traditional Swedish nomenclature
of the months, such as Owl month (March), Grass month (April),
Flower month (May), Hay month (July), Harvest month (August), and
Slaughter month (October). Similar naming patterns regarding farm-
based activities and economy were common in Romania and other
European countries (e.g. Manoliu 1999).

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
6
9

96 Iuga et al.

Seasonality requires detailed scheduling of the various tasks


needed for the different types of land, which strongly influences the
daily life of the whole community. In Romania, the various hay-
making tasks take up more than six months of the year. These tasks
include the clearing of land from stones, anthills, and unwanted
vegetation, spring grazing, mowing the different meadows according
to specific sequences over the summer, occasional autumn grazing,
and, finally in the winter, the transportation of hay to barns. Most
components of this Romanian schedule can be found in Swedish
ethnological source materials (Dahlström et al. 2013).
TEK plays an important role in all the work relating to mowing,
as seen, e.g. in the timing of the activities. In the villages of Maramureș,
as well as in Swedish folklore, people associate the start of mowing
with the ripening of specific plants. In both Romania and Sweden,
this has been connected to Rhinanthus spp: ‘when the clocotici are
heard in the grass, mowing time is here’ (S. C., forty-three years of age,
Șurdești, March 2010). Both the Romanian and Swedish names of this
plant refer to the rattling sound of the ripe seeds. Clocotici or clocotiș
translate as ‘little bell’ in Romanian, and in Swedish höskallra means
‘hay rattle’ and ängsskallra means ‘meadow rattle’.
Another key issue pertains to the community divisions
between ‘work’ and ‘feasts’, where the latter signals the time to rest
and celebrate. The traditions and associated celebrations may signal
the beginning of a new year or another important period of time.
In Sweden, 29 September, the Feast of St. Michael (Mickelsmäss),
marked the end of the working year and the start of a number of
important societal events. Harvesting was so important that lawsuits
could not be called until after the Feast of St. Michael. This was also
the time for markets, and farm workers were given a week’s leave
allowing them to seek new employment, and in the north, it was the
time women returned from their work at the shielings.
Rituals and symbolic actions are still important to Romanian
farmers. Medicinal plants picked on specific feast days are consid-
ered to contain greater healing powers. This applies, for example

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
7
9

Rural Communities & Traditional Ecological Knowledge 97

to 24 June, the midsummer feast associated with the Feast of


St. John, and to 14 September, the Exaltation of the Holy Cross, when
autumn plants such as rose hip are picked (A. M., fifty-six years of
age, Șurdești, April 2010).
In the Romanian tradition, two feasts of the pastoral calendar
are particularly important: St. George’s Day (23 April) is traditionally
celebrated as the beginning of the pastoral year and the warm sea-
son, when sheep are taken to the mountains for their summer graz-
ing. They return at the end of the pastoral summer, on the Feast of
St. Demetrius (26 October) (Nicolau 1998). When Romania switched
from the Julian to the Gregorian calendar in 1924, people continued
going to and from the pastures at the same seasonal time, but now
at different calendar days, sometimes relating to a different feast: St.
George was replaced by Armindeni Day, a celebration held on 1 May,
while St. Demetrius was exchanged for the Orthodox day of the
Archangels Michael and Gabriel (8 November).

4.2.4 Weather Prediction and Tradition


Weather prediction is important in farming communities. TEK
includes numerous experience-based methods of predicting local
weather, centred around winds and clouds, or the most common
local weather developments over the season. The people in Șurdești
(Maramureș) associate southerly summer winds with poor weather,
as they often bring rain. Weather forecasting can also be based on
rituals. One such example is the ‘onion calendar’, used in several
European countries for annual predictions, most commonly for
weather, but also relating to the future health and prosperity of the
farm (Gaignebet 1990: 43–7). On one of the twelve days between
Christmas Day and Epiphany, or sometimes on New Year’s Eve,
an onion is cut in twelve pieces, which are covered in salt and left
overnight in a row on a window ledge. The pieces that dry out by
the following day indicate dry months the following year (Gherman
2002: 58–9). This practice is still widely known, although rarely used.
Similarly, in Sweden, a common belief was that the twelve days

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
8
9

98 Iuga et al.

between Christmas Day and Epiphany could predict the weather of


the twelve coming months (Schön 2005), emphasising the symbolic
importance of the new year.
Even though such predictions are not always correct, their
purpose may be to mentally ‘control’ an unpredictable element, and
to make weather-related decisions collective rather than individual
matters. The purpose may also simply be attempts to influence the
weather by means of ritual and symbolic gestures. Some effects of
bad weather, such as crop failure or dried-out pastures, require fur-
ther action, e.g. alternative uses of the ecosystem or of other sources
of income, which are described in Swedish farming diaries (Westin
et al. 2017). Another alternative is to turn to religious ceremonies, as
portrayed in present-day Romania and in Spanish historical records
(Gomez-Baggethun et al. 2012).
Weather is a central theme in Swedish peasants’ diaries from
the early nineteenth century onwards, both in daily records and in
annual reflections. Many farmers owned thermometers and kept
track of temperature and precipitation, the times when various tasks
were carried out, and the crop and hay yields (Westin et al. 2017).
Comparisons were made with past years and predictions could thus
be produced. In modern Romanian villages, many conversations
between villagers concern the amount of rainfall, seen as the most
important weather factor affecting the crops.
Attempts to influence and appease the weather are also import-
ant within local communities. Their attitudes can be found in cul-
tural knowledge and local heritage. In July, at the height of the
Romanian mowing season, several feast days are considered to be
‘angry with the hay’ (M. C., sixty-five years of age, Șurdești, April
2010), thus, hay working is prohibited (Iuga 2016). Several Romanian
regional legends and celebrations are, for instance, connected to 20
July, the Feast of St. Elijah, seen as the master of thunder. In the vil-
lage of Șurdești, days with similar restrictions related to hay work-
ing are known: the Feast of St. Mary Magdalene (22 July), the Feast
of St. Foca (23 July), the Feast of St. Anne (25 July), and the Feast of

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
9

Rural Communities & Traditional Ecological Knowledge 99

St. Pantaleon (27 July). The stories collected in Șurdești all have in
common that those who do not abide by the restrictions are punished,
usually not directly, but through their haystacks or hay shelters. In
this village, there is a particular place at the foot of the mountain,
known as ‘the petrified haystacks’. Here, two volcanic rocks are asso-
ciated with the legend of a man who built haystacks on a feast day.
St. Peter, who was then wandering the world, saw the man working,
and consequently transformed the haystacks into stone (Figure 4.6).

4.3 Insights from Historical- Ecological


Comparison between Romania and Sweden
Previous sections have illustrated the importance of both current and
recent TEK in northern Romanian rural society, with some Swedish
historical parallels. The comparisons between Romania and Sweden
highlight several aspects of the process where TEK is constantly
changing. This will be discussed in the following sections.
In the past ten years, Romanian and Swedish historians, ecolo-
gists, ethnographers, and anthropologists have collaborated to further
the understanding of the relationships between humans and nature
in rural landscapes (see Wästfelt et al. 2012; Dahlström, Iuga, &
Lennartsson 2013). It is clear that agrarian practices, tools, and
technological and organisational solutions still used in Romanian
mountain villages can be recognised in numerous Swedish written
sources, especially until the mid-nineteenth century. The provision-
ing agrarian ecosystems are also very similar, for example the flora
and vegetation (Dahlström et al. 2013). There are, of course, also
large socio-economic and ecological differences, both in the past and
present, but the similarities can inspire further exploration for how
modern Romania can provide important historical-ecological know-
ledge about Sweden’s agricultural past. Conversely, the agricultural
changes in Sweden can help formulate scenarios for Romania’s future
agriculture. Romania, in particular, can provide insights in TEK con-
cerning the use of semi-natural grasslands (Helldin & Lennartsson
2007; Lennartsson & Helldin 2007). Seen from a European perspective,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
0
1

100 Iuga et al.

Figure 4.6 ‘Petrified haystacks’ in Șurdești village (Romania). The


legend tells that the haystacks were turned into stone by St. Peter, who
encountered a man working with hay on a forbidden day. A black-and-
white version of this figure will appear in some formats. For the colour
version, please refer to the plate section.
Photo: Anamaria Iuga.

Romania has an enormous area (2.4 million hectares) of semi-natural


grasslands, farmed with methods developed through adaptive man-
agement. These grasslands host an exceptional diversity of biological
species, especially in the upland environment (Huband & McCracken
2011; Babai & Molnar 2014).

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
1
0

Rural Communities & Traditional Ecological Knowledge 101

4.3.1 Understanding Swedish Historical Agriculture


and TEK
Our aim is not to ‘export facts’ from present-day Romania that may
apply to historical circumstances in Sweden, but rather to raise
awareness of aspects of daily life that are absent from or hard to
understand through written sources and the current landscape.
Comparisons between the two countries can also reveal previously
hidden or forgotten aspects of Swedish historical ecology that are
worth exploring.
We know, for example, that Swedish rural communities used
to have larger populations, more active landowners, and a higher
degree of collaborative work than today. Since no such communities
are left to study in Sweden, it is difficult to understand daily per-
sonal interaction and its importance for people and for maintaining a
good use of land. In Romanian villages, people can be observed across
the landscape. In summer, families are working in hay meadows and
fields, eating and resting together and talking to others passing by.
When evening comes, men and women continue their discussions on
benches along the main roads.
Although we are aware of the importance of religion in Swedish
rural communities – people were regular churchgoers, holy days were
labour-free, and parish catechetical meetings were important – it is
still difficult to understand the impact of this spiritual world on eve-
ryday life. Experiencing such aspects of life in Romanian villages
raises the awareness of how integrated beliefs, rituals, and other spir-
itual activities can be in daily life, and the ways in which they affect
practical aspects of land use.
Modern Swedish agrarian production is to a large degree pre-
dictable and controllable through the high input of energy, nutrients,
and chemicals. To some extent, TEK and adaptive management are
still present in what is left of Swedish agrarian ecosystems influenced
by natural variation and dynamics. Such low-input arable fields and
unfertilised hay meadows and pastures are considered important
for the preservation of biodiversity and biocultural heritage. Since

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
2
0
1

102 Iuga et al.

almost all grassland-management practices have changed, the effects


of different types of land use cannot be directly studied in Sweden,
but only indirectly studied by combining written sources, cadastral
maps, ecological knowledge, and field visits. In Romania, meadow-
specific management practice and feedback mechanisms between
ecosystem and land, which potentially were the same in Sweden’s
agricultural past, can still be studied in the field. The diverse uses of
hay meadows are further discussed in Chapter 7.

4.3.2 The Present and Future of TEK in Romania


TEK has always been an adaptive process, developing new practices
on one hand and abandoning obsolete ones on the other. Currently,
Romania is facing a situation where TEK, related to low-input agri-
culture, is rapidly disappearing. In Sweden, this process has gone on
for much longer, and some aspects of this development are relevant
for this comparative study.
In many low-producing regions of Sweden, agriculture has
almost completely vanished, which has had profound socio-economic
consequences. Semi-natural grasslands have proven to be an impor-
tant resource for farmers in low-productivity regions, or other areas
where outputs could not compete with others on the market. For
example, beef production based on semi-natural grassland was
important in the Swedish archipelago, and no doubt contributed to
rural development and counteracted migration from rural commu-
nities. Such knowledge can be used to predict and handle effects of
vanishing agriculture in Romania, where the 2.4 million hectares of
semi-natural grassland constitute a tremendous resource.
Consequences of EU rules are also clearly seen, for example
livestock and food hygiene regulations that prevent farmers from
using traditional practices, simply because they do not comply with
EU standards. This could have an enormous impact on land use, as
the authorised solutions require expensive investments incompatible
with traditional small-scale livestock farming. National and EU agri-
cultural policies, although aiming for sustainable development and

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
3
0
1

Rural Communities & Traditional Ecological Knowledge 103

sustainable use of biodiversity, may therefore lead to the loss of TEK,


biodiversity damage, and the counteraction of land-use sustainability.
In Sweden, the total area of grasslands of high natural value
has been reduced to the point that costly and complex restoration is
necessary. It is also difficult to reach environmental goals of habitat
quality, as the required knowledge of how to manage this type of
land may no longer be present. In Romania, no such restoration is yet
necessary, since the areas of managed grassland are still maintained
and the knowledge is intact.

4.4 TEK Devalued and Revalued


The influence on TEK by external factors, such as national policies
altering the conditions for local communities where this knowledge
evolved, has already been discussed. However, TEK itself has also
been and still is the focus of policy and societal interest, with the aim
of either devaluing or revaluing this knowledge base.
Since the mid-nineteenth century, agricultural advice in
Sweden has promoted the development of intensified and rational-
ised production. Local and traditional practices, as well as traditional
breeds and varieties, were undesirable. ‘Older’ practices were, in such
contexts, frequently described as irrational and backwards.
Since Romania joined the EU in 2007, the political framework
of the Common Agricultural Policy (CAP) has had a strong influ-
ence on life in rural communities and the use of TEK. CAP subsid-
ies have become an important source of income for farmers. They
are eligible for agri-environment payment schemes for the manage-
ment of hay meadows and biodiversity-rich pastures, for using trad-
itional agricultural methods, for keeping low-productive livestock,
etc. These EU schemes have complex effects on farming, providing
both support and opportunity, but at the same time restricting some
components of traditional land management (Dunn 2003). Since CAP
contains detailed regulations on how to achieve environmental goals,
some traditional practices are not allowed (Dahlström et al. 2013)
and, as payments and regulations are closely linked, landowners in

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
4
0
1

104 Iuga et al.

mountainous areas are forced to change TEK concerning, e.g. land-


clearing methods, mowing seasons, crop rotation, and multi-annual
cycles, as well as the use of woodland meadows (Stroe 2015; Iancu &
Stroe 2016).
Some traditional practices that policy-makers claim have a
negative impact on the environment are seen by local farmers as
essential for their ecosystem, e.g. the burning of shrubs and old grass
in certain meadow types or temporary cultivation of grass swards.
Such diverging ideas are based on different ways of viewing the land.
On one hand are the sensory perceptions and the traditional criteria
of reading and evaluating territorial practices by those who work a
plot of land. On the other hand are the formalised views of clerks,
informed by digitised data. These differing views may clash during
several stages of the payment cycles. The farmers have little or no
institutional tools to negotiate or correct what they consider to be
erroneous evaluations. In this way, the official interpretation over-
rides local ideas and opinions.
Today, TEK is also recognised internationally, outside the rural
communities where it originates, as part of the intangible heritage of
humanity and a vital component of sustainable development, biodi-
versity, and cultural heritage conservation. International institutions
have contributed to the way we perceive, comprehend, and benefit
from TEK.
The UNESCO Convention for the Safeguarding of the Intangible
Cultural Heritage includes TEK both in its purpose and definitions,
e.g. expressed as ‘practices, representations, expressions, knowledge
and skills’ that can be manifested in social practices, rituals and fes-
tive events, and knowledge and practices concerning nature and the
universe, as well as in traditional craftsmanship (www.unesco.org/
culture/ich/en/convention).
The Convention on Biological Diversity (CBD) contains several
articles relating to traditional knowledge and customary practices
that are relevant for the conservation and sustainable use of biological
diversity. This applies especially to articles 8(j) and 10(c) and Aichi

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
5
0
1

Rural Communities & Traditional Ecological Knowledge 105

biodiversity target 18, which all stress the need for respecting and
maintaining traditional lifestyles in order to preserve, maintain, and
sustainably use biodiversity and cultural landscapes. Implementing
these articles at national, regional, and local levels is a task for each
individual country (e.g. Tunón et al. 2015).
Another reason for the re-evaluation of TEK is that today’s
society appreciates a number of landscape values depending on trad-
itional land use. One such example is biodiversity, as considerable
international interest has been paid to a number of biodiversity-rich
anthropogenic ecosystems, such as semi-natural hay meadows and
pastures (e.g. Gustavsson et al. 2011; Dahlström et al. 2013; Biró
2014; Eriksson et al. 2015). The largest economic incentives to keep
biodiversity-rich grasslands – in Sweden as in most of Europe – are
the payment schemes offered through CAP. The regulations for agri-
environmental support allow and encourage the reintroduction of
necessary traditional practices (Dahlström et al. 2013).
TEK is highly important for the restoration and preservation
of species-rich habitats, as well as the conservation of threatened
species. Conservation practitioners often face difficulties in find-
ing suitable management plans for threatened species and therefore
investigate earlier management practices that supported viable pop-
ulations. Threatened species may depend on their environment in
terms of, e.g. openness, disturbance, and soil type, some of which
require specific land-use practices. These key environmental con-
ditions were in the past created by traditional practices, but today
conservation measures often need to be designed. In order to use his-
torical land use as a template for conservation, TEK of land-use prac-
tices and ecosystems is frequently needed.
In Romania, some traditional customs, forbidden by the com-
munist regime, have been reintroduced (Figure 4.7). Farmers also face
situations where lost knowledge needs reviving for practical reasons.
Biodiversity conservation and CAP can create such incentives, since
it supports farmers who maintain biodiversity and cultural heri-
tage, along with the production of food and fibre. It is also used in

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
6
0
1

106 Iuga et al.

Figure 4.7 The custom celebrates the first man who ploughed his
land, in Șurdești village (Romania). After it was forbidden during the
Communist period, the custom was revitalised in the 1990s. The photo
presents the moment when the celebrated man is taken to the river
and his head bathed in water, in the belief that this would have a good
influence on the crops. Șurdești (Romania), 2012.
Photo: Anamaria Iuga.

other situations: for instance Spanish pastoralists, who traditionally


herded their sheep vast distances on foot, have since the 1940s trans-
ported their sheep by train and subsequently on trucks. As transport
prices have risen with increased oil prices, pastoralists have consid-
ered resuming herding sheep again, thus necessitating a revival of the
transhumance TEK (Oteroz-Rosas et al. 2013).
TEK is also revalued through new communities forming
around the practices themselves (Figure 4.8). Such communities

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
7
0
1

Rural Communities & Traditional Ecological Knowledge 107

Figure 4.8 A traditional wooden fence gärdesgård is under


construction in the archipelago of Stockholm. The knowledge of how
to build the gärdesgård was widespread since the duty and interest to
maintain them was divided among all farmers in the village. Today
fairly few people master these skills in Sweden, and most fences are of
different and more modern types. Hjälmö farm (Sweden), 2014.
Photo: Anna Westin.

share practice and values, but not necessarily rituals and lineage.
They can be seen as ‘communities of practice’ (CoP), i.e. groups of
people who share an interest, a craft, or a profession (Lave & Wenger
1991; Tunón et al. 2014: 43). CoP members are brought together by
joining in common activities, mutual engagement, and their concern
and passion for particular practices. New CoPs have evolved because
of TEK, through people who acknowledge and care for certain heri-
tage values and the impact of specific practices on biodiversity. They
often value self-sufficiency, organic farming, and the use of old know-
ledge in new manners, and are usually part of the so-called neo-rural
or back-to-the-land movements (O’Rourke 1999).
In conclusion, we acknowledge the complexity of fully under-
standing TEK, due to its complex interaction with nature and culture,
individuals and communities, as well as its adaptive and dynamic

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
8
0
1

108 Iuga et al.

nature. However, it is our conviction that a deeper understanding of


the role of TEK in local communities may provide answers to cen-
tral questions in historical ecology, and, conversely, that historical
ecological approaches are essential for the understanding of the pro-
cesses that use, shape, and modify TEK today.

References
Agrawal, A. & Gibson, C. C. (1999). Enchantment and disenchantment: the role
of community in natural resource conservation. World Development, 27 (4),
629–49.
Babai, D. & Molnár, Z. (2014). Small-scale traditional management of highly
species-rich grasslands in the Carpathians. Agriculture, Ecosystem and
Environment, 182, 123–30.
Berkes, F., Colding, J., & Folke, C. (2000). Rediscovery of traditional eco-
logical knowledge as adaptive management. Ecological Applications, 10 (5),
1251–62.
Berkes, F. & Folke, C. (1998). Linking social and ecological systems for resilience
and sustainability. In F. Berkes & C. Folke, eds. Linking Social and Ecological
Systems. Cambridge: Cambridge University Press, pp. 1–25.
Biró, E., Babai, D., Bódis, J., & Molnár, Z. (2014). Lack of knowledge or loss of knowl-
edge? Traditional ecological knowledge of population dynamics of threatened
plant species in East-Central Europe. Journal for Nature Conservation, 22,
318–25.
Dahlström, A., Iuga, A., & Lennartsson, T. (2013). Managing biodiversity rich
hay meadows in the EU: a comparison of Swedish and Romanian grasslands.
Environmental Conservation, 40 (2), 194–205.
Dunn, E. C. (2003). Trojan pig: paradoxes of food safety regulation. Environment
and Planning, 35 (8), 1493–1511.
Einarsson, P. (2015). Traditionell kunskap i modernt lantbruk. CBM:s skriftserie
91. Uppsala: Centrum för Biologisk Mångfald.
Eriksson, O., Bolmgren, K., Westin, A., & Lennartsson, T. (2015). Historic hay
cutting dates from Sweden 1873–1951 and their implications for conserva-
tion management of species-rich meadows. Biological Conservation, 184,
100–7.
Gaignebet, C. (1990). La folle journée de Figaro: entre poils et jardinets. In M.
Mesnil, ed. Les Plantes et les Saisons. Calendriers et Représentations.
Collection Ethnologies d’Europe. Bruxelles: Institut de Sociologie.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
9
0
1

Rural Communities & Traditional Ecological Knowledge 109

Gherman, T. (2002). Meteorologia populară. Observări, credințe și obiceiuri.


Bucarest: Paideia.
Glassie, H. (1995). Tradition. The Journal of American Folklore, 108 (430),
395–412.
Gómez-Baggethun, E., Reyes-Garcia, V., Olsson, P., & Montes, C. (2012).
Traditional ecological knowledge and community resilience to environmental
extremes: a case study in Doñana, SW Spain. Global Environmental Change,
22, 640–50.
Grigg, D. B. (1974). The Agricultural Systems of the World: An Evolutionary
Approach. Cambridge: Cambridge University Press.
Gustavsson, E., Dahlström, A., Emanuelsson, M., Wissman, J., & Lennartsson, T.
(2011). Combining historical and ecological knowledge to optimise biodiver-
sity conservation in semi-natural grasslands. In J. L. Pujol, ed. The Importance
of Biological Interactions in the Study of Biodiversity. InTech.
Handler, R. & Linnekin, J. (1984). Tradition, genuine or spurious. The Journal of
American Folklore, 97 (358), 273–90.
Helldin, J. O. & Lennartsson, T. (2007). Agricultural landscapes in East Europe as
reference areas for Swedish land management. In V. Surd & V. Zotic, eds. Rural
Space and Local Development. Cluj-Napoca: Presa Universitara Clujeana,
pp. 367–70.
Huband, S. & McCracken, D. (2011). Understanding high nature value agriculture
in the Romanian Carpathians: a case study. In B. Knowles, ed. Mountain Hay
Meadows: Hotspots of Biodiversity and Traditional Culture. London: Society
of Biology.
Iancu, B. & Stroe, M. (2016). In search of eligibility: Common Agricultural Policy
and the reconfiguration of hay-meadows management in the Romanian
highlands. MARTOR: The Museum of the Romanian Peasant Anthropology
Review, 21, 129–44.
Iuga, A. (2016). Intangible hay heritage in Șurdești. MARTOR: The Museum of the
Romanian Peasant Anthropology Review, 21, 67–84.
Kideckel, D. (2001). Labor and Society in the Jiu Valley and Fagaras Regions of
Romania. Washington, DC: National Council for Eurasian and East European
Research.
Kideckel, D. (2006). Colectivism și singurătate în satele românești: Țara Oltului
în perioada comunistăși în primii ani după Revoluție. Iași: Polirom.
Kligman, G. & Verdery, K. 2011. Peasants under Siege: The Collectivization of
Romanian Agriculture, 1949–1962. Princeton: Princeton University Press.
Larsson, J. (2009). Fäbodväsendet 1550–1920: ett centralt element i Nordsveriges
jordbrukssystem. Uppsala: Swedish University of Agricultural Sciences.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
0
1

110 Iuga et al.

Lave, J. & Wenger, E. (1991). Situated Learning. Legitimate Peripheral


Participation. Cambridge: Cambridge University Press.
Lennartsson, T. & Helldin, J.-O. (2007). Agricultural landscapes in Eastern Europe
as reference areas for Swedish land management. In K. P. Hasund & J.-O.
Helldin, eds. Valuable Agricultural Landscapes – the Importance of Romania
and Scandinavia for Europe, KSLA 2007: 5. Stockholm: Kungl. Skogs- och
Lantbruksakademien.
Magnusson, L. & Ottosson, L. eds. (2009). The Evolution of Path Dependence.
Cheltenham: Edward Elgar.
Manoliu, V. (1999). Mic dicționar de astrologie și meteorology țărănească.
Bucarest: Mentor.
Mesnil, M. (1997). Etnologul între șarpe și balaur; Mesnil, M. and Popova, A.
Eseuri de mitologie balcanică. Bucharest: Paideia.
Mihăilescu, V. (2000). La maisnie diffuse, du communisme au capitalisme:
Questions et hypothèses, Balkanologie, 4:2. www.balkanologie.revues.org/334
Mihăilescu, V., Nicolau, V., Gheorghiu, M., & Dirnovan, G. (1993). Snagov – trei
proiecții asupra sistematizării. Sociologie Românească, 4 (1), 18–32.
Myrdal, J. (1999). Jordbruket under feodalismen. 1000–1700. Stockholm: Natur
och kultur/ Lts förlag.
Myrdal, J. (2012). Boskapsskötseln under medeltiden. En källpluralistisk studie.
Stockholm: Nordiska Museets förlag.
Myrdal, J. & Morell, M., eds. (2011). The Agrarian History of Sweden. From 4000
BC to AD 2000. Lund: Nordic Academic Press.
Nicolau, I. (1998). Ghidul sărbătorilor românești. Bucharest: Humanitas.
O’Rourke, E. (1999). Changing identities, changing landscapes: human-land rela-
tions in transition in the Aspre, Roussillon. Cultural Geographies, 6 (1), 29.
Österling, P.-A. (2010). Dialekt och folkminnesarkivens material – etnologi på
Institutet för språk och folkminnen (SOFI) exemplet ULMA. In H. Tunón and
A. Dahlström, eds., Nycklar till kunskap. Om människans bruk av naturen.
Uppsala: Centrum för Biologisk Mångfald and Stockholm: Kungl. skogs- och
lantbruksakademien, pp. 59–66.
Oteros-Rozas, E., Ontillera-Sánchez, R., Sanosa, P., Gómez-Baggethun, E., Reyes-
García, V., & González, J. A. (2013). Traditional ecological knowledge among
transhumant pastoralists in Mediterranean Spain. Ecology and Society, 18
(3), 33.
Persson, J. & Nilsson, N. Ö. (1996). Lien och dess marker. Stockholm: Lts förlag.
Ruiz-Mallén, I. & Corbera, E. (2013). Community-based conservation and tra-
ditional ecological knowledge: Implications for social-ecological resilience.
Ecology and Society, 18(4), 12. http://dx.doi. org/10.5751/ES-05867-180412

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
1

Rural Communities & Traditional Ecological Knowledge 111

Schön, E. (2005). Folktrons ABC. Stockholm: Carlsson.


Stahl, H. H. (1946). Sociologia satului devălmaș românesc. Organizarea economică
și juridică a trupurilor de moșie. Bucarest: Fundația Regele Mihai I.
Stroe, M. (2015). Rigid norms, flexible subjects: socio-ecological resilience in con-
tested hay landscapes. In G. Troc & B. Iancu, eds., Modes of Appropriation and
Social Resistance, Bucharest: Tritonic, pp. 103–20.
Szelenyi, I., ed. (1998). Privatizing the Land. Rural Political Economy in Post-
Communist and Socialist Societies. London: Routledge.
Torsello, D. (2003). Trust, Property and Social Change in a Southern Slovakian
Village. Munster: Lit. Verlag.
Tunón, H. (2015). Studiet av bruket av naturen. Ur etnobiologins historia. In H.
Tunón, ed., Etnobotanik. Planter i skikog brug, i historien og i folkemedicinen.
Vagn J. Brøndegaards biografi, bibliografi og artikler på dansk i udvalg. Vol.
1. Stockholm: Kungl. Skogs- och Lantbruksakademien and Uppsala: Centrum
för Biologisk Mångfald, pp. 33–46.
Tunón, H., Kvarnström, M., Axelsson Linkowski, W., & Westin, A. (2014). Hur
bör Sverige genomföra artiklarna 8j and 10c i syfte att uppnå Aichi-mål 18
i FN:s Konvention om biologisk mångfald?. Uppsala: Centrum för Biologisk
Mångfald.
Verdery, K. (2003). The Vanishing Hectare: Property and Value in Postsocialist
Transylvania. London: Cornell University Press
Wästfelt, A., Saltzman, K., Gräslund Berg, E., & Dahlberg, A. (2012). Landscape
care paradoxes: Swedish landscape care arrangements in a European context.
Geoforum, 43, 1171–81.
Westin, A., Isacson, M., & Lennartsson, T. (2017). Land and labour as resources of
an integrated peasant economy in a Swedish mining district during the 1860s
great famine. In A. Panjek J. Larsson and Mocarelli. L, eds., Integrated Peasant
Economy in a Comparative Perspective – Alps, Scandinavia and Beyond,
Koper: University of Primorska Press, pp. 309–32.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:32, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.004
2
1

5 Baselines and the Shifting


Baseline Syndrome – Exploring
Frames of Reference in Nature
Conservation
Tuija Hilding-Rydevik, Jon Moen, and
Carina Green

5.1 Introduction
The conditions of most current landscapes are shaped by historical,
cultural, and social human activities interacting with biological,
chemical, and geological processes. Nature conservation research and
management practices are still in the process of understanding this
‘socio-natural hybridity’ (Rice 2013) of landscapes and of our environ-
ments. This reflects a need for increased historical and multidiscipli-
nary awareness in nature conservation research, as well as in policy
and management practices in order to fully appreciate the history
and complexity of human-environmental interaction and hybridity.
A historical and multidisciplinary awareness also provides research
disciplines, individuals, and societies with a broader understanding
of how and why we are who we are, and why we think and act as we
do. This approach is important for critical self-reflection and crucial
for steering practice in more sustainable directions, be it in research
or in wider contexts.
The aim of this chapter is to discuss the challenges and con-
tributions of a widened historical-ecological frame of reference in
the context of conservation research and practice. A frame of refer-
ence is here used as an umbrella concept meaning ‘a set of ideas,
conditions, or assumptions that determine how something will be
approached, perceived, or understood’ (www.merriam-webster.com).
We explore frames of reference from two aspects. The first is the role
of a clear historical awareness in the process of choosing baselines
112

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
3
1

Baselines and the Shifting Baseline Syndrome 113

for conservation. Baselines are fundamental for conservation policy,


planning, and management. For example, they can represent the con-
servation target, i.e. an ‘ideal state’ to strive for, a reference ecological
community, or the chemical state of freshwater, and can later be the
foundation for evaluation of conservation measures. We argue that
baselines are always socially constructed,1 and thus influenced by
political agendas, economic realities, preconceived ideas, and socio-
cultural ‘understandings’ of human impacts on nature. This is obvi-
ous when working with conservation in cultural landscapes where
baselines are influenced by the combined choices of time periods,
land-use regimes, species groups, etc., as discussed in Chapter 9. In
particular, baselines aimed at reconstructing ‘pre-human’ or ‘natural’
states are products of specific socio-cultural presumptions, and the
following discussion focuses on these ‘pristine nature baselines’.
The second aspect is the shifting baseline syndrome (SBS),
described as ‘inter-generational changes in perception of the state of
the environment’ (Sáenz-Arroyo et al. 2005: 1557). We broaden the
frame of reference for SBS research by relating it to another, very
relevant research field, that of social memory. The use of the term
syndrome should be questioned, since it suggests at least a degree of
abnormality when describing the shifts and changes in what people
and generations remember, e.g. in terms of fish catches and biodiver-
sity. Memory research suggests that concepts such as baselines and
SBS do not capture the complexities of the ‘memoryscapes’ imbed-
ded in all human societies. It also proposes that baselines and SBS

1
Social constructionism is a main theoretical orientation. In this orientation,
human beings are seen as social and our perception of reality is socially
constructed in communication and interaction between people. Taken-for-granted
knowledge is criticised; knowledge is historically and culturally dependent
and is constructed and reconstructed in social processes. Language is seen as a
precondition for thought and is a form of social action (Berger & Luckmann 1979;
Burr 1995; Gergen 2001). By arguing that baselines are always socially constructed,
we thus mean that there exist no self-evident ‘objective’ baselines. A choice of
baseline is always a value choice and thus dependent on current world views,
history, policy, research traditions, availability of data, etc. These are conditions
humans create through their social relations and through their interaction with
the environment and with things.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
4
1

114 Hilding-Rydevik et al.

are results of the specific cultural context from which the scientific
activity has sprung.
The problems discussed in relation to baselines as SBS bring
to light the consequences of the nature-culture divide in natural and
social sciences and the humanities (for a discussion on the nature-
humans dichotomy, see e.g. Latour 2004). We conclude by argu-
ing for an increased historical and multidisciplinary awareness in
nature conservation research and the need for new integrative con-
cepts going beyond the nature-culture divide. Historical ecology is
well suited to examine the complex relationships between human
activities and the environment, although currently seldom used in
baseline and SBS discussions in conservation (De Vries 2005). In
this chapter, the concepts of baseline and SBS are used in order to
highlight:

1. The profundity of historical interaction between nature and human


culture, and the problems that may arise if this is not recognised when
baselines are investigated and decided on.
2. How research from the humanities and social sciences can contribute to
research related to baselines and shifting baseline syndrome in nature
conservation.

5.2 Baselines as a Concept in Nature


Conservation
Baselines are used in many parts of society and research, such as
health care, business, and nature conservation. A baseline may be
defined as ‘a line serving as a basis; especially: one of known meas-
ure or position used (as in surveying or navigation) to calculate or
locate something’ (www.merriam-webster.com). In nature conserva-
tion and environmental management, baselines have been discussed
mainly in marine sciences, usually in relation to the state prior to
large-scale human exploitation of marine resources (Pauly 1995;
Campbell et al. 2009; Jackson, Alexander, & Sala 2011). The concept
is defined and used in a number of different ways. Maron and col-
leagues (2015: 505) give the following examples:

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
5
1

Baselines and the Shifting Baseline Syndrome 115

Baseline: a state or trajectory (e.g. of a system) used as a comparator.


The term is used in many contexts: shifting baseline is often used
to refer to the way in which our concept of what is ‘normal’ or ‘pre-
impact’ changes over time (e.g. Papworth et al. 2009; Pauly 1995);
baseline data often refers to data that reflect the starting-point or
reference state of a system before some expected perturbation (or
restoration) (e.g. Downs et al. 2011); a baseline site can refer to
a ‘control’ or ‘reference’ site in an experimental design or natu-
ral experiment (e.g. Brinck and Frost, 2009; Golding et al. 1997);
a baseline trajectory is a counterfactual trajectory that describes
how a system is expected to behave in the absence of some pertur-
bation or action (e.g. Costa et al. 2000).

Baselines have been proposed for, e.g. deep-sea conservation in the


North Atlantic (Foster, Foggo, & Howell 2013), in relation to large
ungulate conservation in Asia (Gray et al. 2012), in marine spe-
cies conservation (McClenachan, Feretti, & Baum 2012), and the
conservation of Yellow Crazy ants in Samoa (Hoffmann, Auina, &
Stanley 2014). Baselines are also set in relation to the Convention on
Biodiversity (CBD 2007), and in relation to the emission level targets
of the International Panel for Climate Change (IPCC).
In environmental and nature conservation policy and manage-
ment, different goals are set by, e.g. researchers, policy-makers, and
NGOs in order to guide the formulation of legislation, guidelines,
and practical measures. Baselines are set in order to establish goals,
to outline what is worth saving or restoring in conservation, or to
measure against what to value conservation interventions. A current
example is biodiversity offsets, ‘compensating for losses of biodi-
versity at an impact site by generating ecologically equivalent gains
elsewhere’ (Maron et al. 2012: 141; see also Bull et al. 2016). These
often aim to prevent net loss of biodiversity, although this raises the
question: no net loss compared to what? (Maron et al. 2015). In such
contexts, baseline assumptions are very important, as Bull and col-
leagues (2014) demonstrate, since the stated success of a biodiversity

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
6
1

116 Hilding-Rydevik et al.

offset action in relation to the ‘no net loss’ goal is heavily dependent
on the choice of baseline.
The way in which the goals are set may be based on choices
of, for example, temporal, geographical, or historical reference points
or periods. Ferraro and Pattanyak (2006) and McDonald-Madden
and colleagues (2009) (in Bull et al. 2014: 800), however, state that
‘appropriate frames of reference are not widely used’ and that this ‘is
a problem for contemporary conservation’. Bull and colleagues (2014:
800) also point out that the choice of frame of reference ‘is a critical
component of the process of conservation’ and that this choice, in
conservation policy, often is ‘unstated and implicit’. With the aid of
historical ecology, more appropriate baselines for conservation can be
determined, which (1) provide a wider understanding of the historical
and human-environment-interaction processes that have shaped our
landscapes; and (2) acknowledge such choices as ‘value judgements’
that need to be transparent and explicit.

5.2.1 Historical and Biocultural Awareness


Humans have undoubtedly had a profound influence on the global
environment (e.g. Cardinale et al. 2012; Hooper et al. 2012). Scientific
answers and other assumptions concerning when, where, and to
what extent this happened have considerable effect on the choice of
baselines in conservation research, practice, and policy. How such
baselines are set in conservation is sometimes rather not considered
regarding human impact on nature in the past. This applies to con-
servation in cultural landscapes (see Chapter 9), and especially to
baselines aiming for an environmental status that appear to mirror
nature before it was altered by humans, or to states existing prior to
human exploitation (Campbell et al. 2009). Such baselines have been
used ever since the beginning of nature conservation, and have been
subject to research. Hunter (1996: 695), for instance, discusses how
to define ‘natural’ and suggests that by using it, ‘to mean “without
human influence” in the context of conservation would help define
clear benchmarks for managing ecosystems both inside and outside

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
7
1

Baselines and the Shifting Baseline Syndrome 117

ecological reserves’, and Knowlton and Jackson (2008: 216) discuss


the need for ‘understanding how pristine ecosystems functioned
before human disturbance’.
Using such ‘pristine environments’ as benchmarks for conser-
vation baselines causes a number of problems, normative as well as
scientific and practical. Some of these problems will be discussed
later in this chapter, and it is concluded that it is difficult to describe
the baseline accurately, agree on the description, and restore the eco-
system to the desired state (cf. Campbell et al. 2009).
Human influence goes so far back in time that geologists
are identifying the current geological epoch as the ‘Anthropocene’
(e.g. Lewis & Maslin 2015). The term Anthropocene acknowledges
humans as important geological and environmental agents. This has
caused many critical elements of the bio-geo-physical environment
to exceed their Holocene ranges, including greenhouse gas concen-
trations, ocean acidity, nitrogen cycles, mineral transport and ero-
sion, and human appropriation of net primary production, as well as
extinction and invasion rates (Corlett 2015). The start date for this
epoch is hotly debated. Dates range from 1.8 million years ago (when
fire was first used to modify ecosystems), through 15,000 years ago
(megafauna extinctions), to 7,000 years ago (the rise of agriculture)
(see references in Corlett 2015 and Lewis & Maslin 2015). Lewis and
Maslin (2015) have also suggested two even more recent dates: AD
1610 based on a minimum level of atmospheric CO2 due to popula-
tion declines in the Americas after colonisation by Europeans, and AD
1964, based on a peak in 14C after the detonation of nuclear bombs.
Even though humans have been affecting global processes for a
very long time, local effects vary considerably. This poses problems
when investigating which areas are affected by humans and to what
extent. Areas with intense agriculture are more altered than marginal
areas, and a lot of research has been focussed on the spread of agri-
culture and the subsequent effects on the biosphere. Ellis (2011), for
instance, describes anthropogenic transformations of the terrestrial
ecosystems from 6000 BC to AD 2000. In his maps for AD 2000,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
8
1

118 Hilding-Rydevik et al.

large parts of the boreal and arctic biomes, together with parts of
the Amazon, are described as wildlands, i.e. areas ‘without human
populations or use of land’. (Ellis 2011: 1014, figure 2). Even these
areas may, however, show effects of long-term human use, albeit
subtler and overlooked. This is similar to Denevan’s (1992) ‘pristine
myth’, describing how the landscape of the Americas prior to 1492
was perceived. So the ‘myth’ actually refers to historical landscapes
that were biocultural, i.e. altered by indigenous populations for thou-
sands of years, and not really pristine. Subtle traces of human pres-
ence cause problems in landscape history studies, and landscapes
that are misinterpreted as not having histories of human use have
critical implications for conservation.
Baffin Island in northern Canada is an example from the high
Arctic which illustrates the importance of historical awareness when
deciding what is and what is not pristine. The island has been inhab-
ited by different groups (Inuit, Palaeo-Eskimos) since around 2500
BC (Michelutti et al. 2013). In one study, the long-term impacts of
hunting and butchering marine mammals in freshwater ponds were
studied using palaeolimnological methods. Human activities left dis-
tinct geochemical signals in the sediments which in turn affected
diatom communities. The cold temperatures cause a slow rate of
decomposition of bones of butchered animals, subsequently affecting
the ponds for long periods of time. The signals in ponds close to set-
tlements were found to be distinct from those further afield with no
or only periodic human use. Different ponds in the area thus provided
different indications of ‘pristine-ness’, and consequently using either
affected or unaffected ponds as bio-geo-chemical baselines in a con-
servation setting would give very different results.
Many of the ecosystems valued in conservation are as much
cultural as ecological landscapes (Lorimer et al. 2015). Nonetheless,
there is limited knowledge of past land use in many areas protected
for their nature values (Setten & Austrheim 2012; Natlandsmyr &
Hjelle 2016). One example is Tjeggelvas, a nature reserve in northern
Sweden, considered one of the few pristine boreal forests in northern

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
9
1

Baselines and the Shifting Baseline Syndrome 119

Figure 5.1 Photo of a Sami family, probably from the late nineteenth
century. The picture is taken in the central part of what is now the
Tjeggelvas nature reserve and is currently considered an example of a
pristine boreal forest, i.e., untouched by humans.
Photo from the photo archive at the Silver Museum, Arjeplog.
Photographer unknown.

Sweden (Josefsson 2009; Josefsson, Hörnberg, & Östlund 2009). The


area has, however, been used by the indigenous Sami people for cen-
turies and subsequently shows subtle traces of human impact. In the
past, from about 1700 to 1900, the area included at least two Sami
taxation lands, where Sami households practised hunting, fishing,
and reindeer husbandry (Figure 5.1; Josefsson, Bergman, & Östlund
2010). Apart from various historical and archaeological remains, such
as hearths, storage facilities, bark peeling, and trail markers (Rautio,
Josefsson, & Östlund 2014), changes in the vegetation can also be
detected. Close to the areas that have been used the most, the reserve
forest has a higher proportion of deciduous trees, and the field layer
vegetation a higher proportion of grasses and forbs (Josefsson et al.
2010). Such land-use traces do not detract from the old-growth char-
acteristics and high conservation values of the reserve. By applying

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
0
2
1

120 Hilding-Rydevik et al.

an uncritical ‘pristine’ label, however, there is a potential risk of


setting a conservation baseline which overlooks anthropogenic dis-
turbances on forest and vegetation, essential for maintaining conser-
vation values of the area. In the Tjeggelvas case, it is obvious that
the local Sami reindeer herding culture has had a direct influence on
the vegetation and biodiversity of the area. This should be evaluated
when setting baselines, both for ecological reasons and because the
impact may be considered part of a biocultural heritage adding to the
conservation value of the site. The example also shows an inclin-
ation, common in many instances throughout colonial history, of
viewing indigenous peoples as so directly associated with their lands
that they complete, rather than have a real impact on them (Sörlin
1991: 266).
The ideology of naturalness, with its focus on pristine environ-
ments, has its roots in the long-standing Western idea of a nature-
culture dualism or divide (e.g. Gómez-Pompa & Kaus 1992; Escobar
1999; Hornborg & Pálsson 2000). This view equates ‘culture’ with
all human artefacts, and ‘nature’ with an environment separate from
humans, as two separate realms of reality. A contrasting view, based
on Anthropocene research, is that it is not possible to distinguish
between ‘nature by herself as a standard’ and ‘nature modified and
polluted by humans’ (Haila 2000: 158). Haila (1997) and Comer (1997)
criticise Hunter’s (1996) use of ‘natural’ to mean ‘without human
influence’, and argue that since humans are creatures of nature, and
all human activities are based on processes of nature, it is impossible
to separate us from it.

5.2.2 Processes for Setting Baselines


Haila (1997) points out that what is regarded as ‘nature’ and ‘natural’
is culturally laden. Any associated management decision will include
conflicting values and interests that need to be handled through
deliberative processes. This idea includes an acceptance that there
is no objectivity in the process of choosing baselines. These choices
are always value judgements (Lélé & Norgaard 1996; Campbell et al.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
1
2

Baselines and the Shifting Baseline Syndrome 121

2009; Bull et al. 2014) and products of our frames of reference. These
judgements need to be addressed explicitly and include different
aspects (scientific, geographical, cultural, etc.) in a transparent pro-
cess. Questions such as what is worth preserving, why, for whom,
and how, as well as how, when, and where to use baselines, must be
determined and should be decided by using participatory and demo-
cratic methods that include local inhabitants, governing bodies,
archaeologists, and other knowledge communities.
This need was obvious in the case of a conservation project in
Tanzania which focussed on eliminating environmentally destruc-
tive practices in a forest area. A baseline was set to determine the
effects of the proposed actions (De Vries 2005). This was, however,
set without an understanding of the locals’ historical use and role of,
as well as their ties to, the forest, and their historically based distrust
of outsiders. Since the conservation actions proposed (e.g. reduced
access to the forest) appeared to be based on a classical colonial
view of a tribe in a wilderness landscape, they were resented by the
local population. De Vries (2005) argues that an alternative approach
would have been to develop a baseline based on ecological data, cou-
pled with information on local history and cultural memory on mul-
tiple societal levels in a careful political facilitation process aimed at
reaching a public compromise: a ‘democratisation of the baseline’.
Participatory methods and actions have increasingly been
sought in environmental contexts as these are often complex, uncer-
tain, and involve many scales (e.g. local, regional, national, global).
With a deep disagreement on values added to a high degree of sci-
entific uncertainty, the situation is often referred to as a ‘wicked
problem’ (Balint et al. 2011). There is some evidence that stakeholder
participation can enhance the quality of environmental decisions
(see Reed 2008 for a review). It is however important that issues of
power, equity, trust, and learning are addressed in the processes (e.g.
Cook & Kothari 2001).
A huge number of collaborative and participatory methods
exist, and it is beyond this chapter to review them all. What follows,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
2
1

122 Hilding-Rydevik et al.

however, is an example that illustrates the complexities of public par-


ticipation when setting a baseline (Valinia et al. 2012). The European
Union Water Framework Directive (WFD; EU Directive 2000) sets
goals for watersheds and water management. The overall WFD
management goal for a good ecological status of a body of water is
dependent on establishing ‘reference conditions’, or baselines. This
includes that there are to be ‘no, or very minor, anthropogenic altera-
tions of the physico-chemical and hydromorphological quality ele-
ments of the water body’ (EU Directive 2000: 38), i.e. in essence that
it is ‘pristine’, or in some aspects ‘natural’ (Bishop et al. 2009). This is
problematic since: i) most European waters have been subject to some
degree of human alteration; ii) there is no guidance on how to distin-
guish between no and very minor human alterations; iii) the criteria
for choosing ‘reference conditions’ are subject to different interpreta-
tions; and iv) an undisturbed state might not be an appropriate goal
for water management (Valinia et al. 2012). In recognition of some of
these difficulties, the WFD also encourages public participation when
setting up goals for a good ecological status. Valinia and colleagues
(2012) initiated a case study on the status of Lake Rotehogstjärnen
in Sweden involving both landowners and researchers, where they
argued that the inhabitants possess local knowledge which could be
used to conceptualise reference conditions (cf. Dallimer et al. 2009).
Valinia and colleagues (2012) further argued that a combination of
local knowledge and monitored data was important for recognising
alternative reference conditions. Valinia and colleagues (2012) thus
argued for a reflective process of defining a reference condition.

5.3 The Shifting Baseline Syndrome


An awareness of historical ecology and of human-environment
interactions is needed for conservation research and policy in
order to include all complexities in human-environment relation-
ships when setting baselines. The remaining part of this chapter
addresses a specific aspect of historical awareness, that which is
based on memory.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
3
2
1

Baselines and the Shifting Baseline Syndrome 123

SBS has received increasing attention within conservation


biology since Pauly coined the term (1995). The syndrome, as stated
initially in this chapter, describes the generational changes in how
the state of the environment is perceived. This may result in a situ-
ation where ‘the modified ecosystem state comes to be perceived as
the norm and the historical ecosystem state is forgotten’ (Brown &
Trebilco 2014: 684). Ortmann (2010) considers shifting baselines an
example of ‘drift’, where the slowness and imperceptibility of gradual
changes extend beyond human time frames (e.g. a generation). When
SBS is used in nature conservation, it is based on the assumption
that humans, as individuals and collectives, are able to accurately
describe some previous baseline through their memories of the past.
The term shifting baseline syndrome assumes the existence of a
memory ‘shift’, labelled a ‘syndrome’, implying that it is not a nor-
mal process in peoples’ lives and in society.
Many authors have described the risks involved in omitting
historical data in relation to SBS. This could involve, e.g. that invad-
ing species over time come to be regarded as native (Clavero 2014);
that people may forget how vibrant the coral reefs once were, lead-
ing to less ambitious conservation goals (Drew, Philipp, & Westneat
2013); or that people are no longer capable to interpret the sights and
sounds of animals in the forest that their parents could, leading to a
decline in knowledge of its species (Kai et al. 2014). This could result
in, for example, assessments of conservation status that are too opti-
mistic, recovery targets that are set too low, and fisheries quotas that
are set too high (McClenachan et al. 2012).
Some nature conservation research has examined the under-
standing of how such perceptions change. Papworth and colleagues
(2009), for instance, discussed various possible processes involved in
baseline-perception changes. Implicit to the SBS is that the environ-
ment actually has changed. If this is the case, generational amnesia
may be occurring if there are age- or experience-related differences in
perception. If no such differences are found, but individuals believe
that current conditions also occurred in the past, it would be an

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
4
2
1

124 Hilding-Rydevik et al.

indication of ‘personal amnesia’. Papworth and colleagues (2009)


also identified two other combinations of environmental change and
observer perceptions that are usually not included in definitions of
SBS: ‘memory illusion’ (i.e. to perceive that the environment has
changed when in fact it has not) and ‘change blindness’ (i.e. to per-
ceive that the environment has not changed when in fact it has). Daw
(2010), for instance, describes how memory illusion may develop
within local fisheries as exaggerated memories of extreme catches
over time become perceived as trends. Another term used for the loss
of understanding of earlier ecosystem states is the ‘cultural amnesia’
offered by Drew and colleagues (2013).
Dallimer and colleagues (2009) investigated changes to agricul-
ture and habitat in the Peak district in northern England, comparing
maps and census data with information provided at local stakeholder
workshops. Their results show that some major historical changes,
such as an intensification of sheep farming, corresponded with the
perceptions of the stakeholders, while other changes, such as trends
in agricultural labour and the dynamics of vegetation changes, did
not. They attribute the differences between the historical data and
people’s memories to perceptions of scale. They argue that stakehold-
ers tend to have holistic memories, making it difficult to distinguish
complex drivers and patterns. Additionally, data and stakeholder per-
ceptions may also differ because of different spatial scales. Since local
patterns differ, stakeholder perceptions can vary in accordance with
specific local knowledge, while historical data, which are aggregated
on broader scales, fail to display such local variation. Since stake-
holder perceptions of change do not always correspond to recorded
history, the authors question the value of stakeholder opinions and
workshops. Instead they suggest that patterns of historical changes
should be based on long-term ecological data. This means that they
contradict the implicit assumption that stakeholders have some kind
of objective and ‘accurate’ memory. Papworth and colleagues (2009)
also argue that local ecological knowledge should be used with caution
because of the risk of SBS. We argue that neither the local knowledge

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
5
2
1

Baselines and the Shifting Baseline Syndrome 125

nor the aggregated data are ‘wrong’, but are instead representations of
different temporal and spatial scales. It ought to be possible to merge
these differences in deliberative processes, such as the one described
earlier in this chapter (Valinia et al. 2012). Studies in historical ecol-
ogy show also that ‘people’s perceptions of species abundance from
historical narratives are generally consistent and useful as starting
points for ecological studies. This was demonstrated when using an
innovative method for testing the validity of interpretations of peo-
ple’s anecdotes on species abundance (Al-Abdulrazzak et al. 2012).
SBS is often invoked as a problem for conservation, although
this is seldom supported by empirical evidence (Papworth et al.
2009). We agree with Jackson and Alexander (2011), who point out
that the SBS phenomenon requires an interdisciplinary approach in
order to be able to contribute to the understanding of past changes in
social, historical, and scientific contexts. One way forward would be
to include research on memory processes.
As outlined previously in this chapter, individual and some
type of collective memory are important ingredients in the SBS dis-
cussion. The concept has mainly been developed in a conservation
research context. What memory is, however, and how it is created,
both individually and in society, is rarely explicitly stated in this dis-
cussion. Assumptions and concepts related to SBS are found in the
large body of interdisciplinary memory research established as early
as the beginning of the twentieth century. These could provide valu-
able ideas for SBS research, e.g. in relation to a more elaborate set of
concepts based on a rich amount of empirical results. The aim of the
following sections is to describe some of these relevant results and to
discuss SBS in relation to long-standing memory research.

5.3.1 Memory as Something We Do


The concept of memory and loss of memory have a lot to contrib-
ute to the SBS discussion. To gather memories in local communi-
ties, from individuals and groups, is a common way of producing data
about historical states of the environment (size of catches, abundance

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
6
2
1

126 Hilding-Rydevik et al.

of different species, etc.). SBS researchers label changes in local stake-


holders’ perceptions of the past as various types of memory loss (e.g.
personal amnesia, memory illusion, and change blindness), when
such changes do not correspond to scientific change data (assumed to
be more objective).This would indicate that humans could have ‘cor-
rect’ memories. Memory research, however, unveils a much more
complex picture of memory and the different processes that shape
both collective and individual memories.
Memory studies are interdisciplinary and the main contribu-
tions come from the humanities and social sciences (psychology,
sociology, history, anthropology, etc.). Various concepts of memory
studies are also used in environmental and nature conservation
research, although not related to the SBS discussion, e.g. social-
ecological memory (Barthel, Folke, & Colding 2010; Nykvist &
Heland 2014) and botanical memory (Ryan 2013). In nature conser-
vation, memories and remembering are also discussed in relation to
resilience studies, where they may be essential for ecosystem man-
agement (Barthel et al. 2010). Nykvist and Heland (2014) also show
how memories of past natural resource use may be important for the
formation of group identities. They show how an unwillingness to
abandon historical traditional resource use prevented the adaptation
to climate change and the use of new alternative resources. This
could be seen as an example of how memories linked to group iden-
tity may result in less resilient behaviour.
The concept of memory is not easy to define or describe, as
illustrated by Olick, Vinitzky-Seroussi, and Levy (2014: 134): ‘it is
not clear that even in psychology there is such a thing as “memory”
per se.’ Tulving refers to approximately twenty-five different kinds
of memory and Roediger and colleagues (2002) identify 256 (in Olick
et al. 2014). Roediger and colleagues (in Olick et al. 2014: 134) go on
to state that: ‘Memory is a single term, but refers to a multitude of
human capacities,’ and, ‘the single term memory does not do justice
to the underlying concepts it represents.’ This is reflected in memory
research, where many different concepts of memory are used.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
7
2
1

Baselines and the Shifting Baseline Syndrome 127

The concept of SBS mainly relates to memory as ‘something


we have’. The memory of and remembering, however (e.g. the abun-
dance of species or a size of harvest), are part of a much more com-
plex process. Remembering deals not merely with what our brains
can objectively recall from observations and personal experiences.
Memory research clearly states that memory should be seen more as
‘something we do rather than something we have’ (Terry 2013: 476).
This implies that memories are constructed, and remembering takes
place through complex cognitive, social, and societal processes of
which we, as humans, are a part. Memories have an impact on how
we interpret what we experience through our senses (Zelizer 1995 in
Olick & Robbins 1998; Rothberg 2009 in Healy & Tumarkin 2011).
When working with locally sourced information on past and present
states of resources, on which local stakeholders rely, it is import-
ant to understand that images of the past may change over time and
that groups may use the past to express particular interests (Olick &
Robbins 1998).

5.3.2 Individual and Collective Memory


Concepts such as generational and cultural amnesia, as used in SBS
discussions, refer to forms of collective memory, but without dis-
cussing whether this really exists. Gavriely-Nuri (2014) argues that
the concept is often used in a metaphorical sense. Our interpretation
of the use of SBS-related concepts, such as generational and cultural
amnesia, is that they are metaphors for attempts to find generalised
concepts to describe the memory loss of individuals and groups in
terms of how species and ecosystems disappear and change.
The concept of collective memory has developed into a
research field since Hugo von Hofmannsthal first explicitly used
it in 1902 (Olick & Robbins 1998). Collective memories can be
described as shared by a group, regardless of whether individual
members have personal experiences of them. Central to this is how
narratives of historical events and conditions are acknowledged
and retold. It is important, however, to recognise the relational and

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
8
2
1

128 Hilding-Rydevik et al.

changing aspects of collective memories. They are not exact recol-


lections or uncontested reproductions of actual events, but rather
representations of the past that also give meaning to the present.
Hence, they have been described more as indicators of how the pre-
sent is perceived (Kenny 1999: 437). This does, however, not mean
that these memories can be discarded as ‘false’ accounts of the past.
In a sense, collective memories can be seen as representing the
meaning of the past, rather than mere objective facts, e.g. in relation
to the abundance of species and size of fish catches, or the success
of game hunting.
Collective memories can be described as contexts where per-
sonal experiences can be shaped and articulated (Linde 1997, 2000).
But it is only individuals who remember: ‘Individual memory is a con-
crete memory: the memory of things or events that one has personally
experienced, which can be remembered because the person was pre-
sent when the thing appeared or the event unfolded’ (Gavriely-Nuri
2014: 47 summarizing Halbwachs 1992). ‘CM [collective memory],
in contrast, extends the scope of individual memory by incorporat-
ing information that goes beyond one’s own experience of the world’
(Wilson 2005 in Gavriely-Nuri 2014: 47).
Many other concepts are used and discussed in memory
research. Some of these are used in SBS research, but without recog-
nising the different meanings that can be applied. One such concept is
the term cultural memory, as outlined by Terry (2013: 475): ‘although
used variously, [cultural memory] suggests a sense of memory as
connected to sociocultural contexts, and my focus is on the order of
memory that sees social groups constructing a shared past via media,
institutions, and practices, this in turn contributing to the shaping of
communities.’ In SBS research, it would also be useful to differenti-
ate among various kinds of memory, such as ‘episodic’ and ‘seman-
tic’ memory (Gavriely-Nuri 2014). In SBS research, much attention is
given to episodic memory, i.e. what you have personally experienced
and felt, e.g. different levels of fish catches or biodiversity. The lesser-
used semantic memory is based on general facts and includes general

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
9
2
1

Baselines and the Shifting Baseline Syndrome 129

and structured knowledge which is not related to a specific event


(Gavriely-Nuri 2014).
An acknowledgement of the ongoing empirical discussion
within memory research as a whole would benefit the SBS discus-
sion and help it become more nuanced. According to Belvedresi
(2014: 109): ‘The first [aspect] is related to the concept of “collective
memory”: what phenomenon of the social world it can describe, how
to determine whose memory it is, what actors produce it (and repro-
duce it), how to determine its duration, its material substrata, and so
on; we might speak of the “empirical” problems associated with the
identification of certain aspects of shared life.’
Such ‘empirical problems’ are prevalent in the SBS research, and
mostly deal with episodic memory issues (Gavriely-Nuri 2014), i.e.
stakeholders with individual and personal memories of past events,
and those directly engaged in the management of natural resources.
The memories of this group are also called ‘social-ecological mem-
ory’ (Barthel et al. 2010). Social-ecological memory is defined as the
accumulated experiences and history of ecosystem management
collectively held by a community in a social-ecological system, i.e.
groups that have episodic memory. Not included in SBS, however,
are all the other groups in society that are involved in remember-
ing individually and collectively, e.g. politicians, public conservation
officials, NGOs, and, not least, conservation researchers.

5.3.3 Mnemonic Practices


Memory research also deals with constructing mnemonic practices
and formalising memory institutions for the future. Mnemonic prac-
tices refer to tools used for remembering, ranging from the old art of
storytelling to modern electronic means of recording and transmit-
ting information (see e.g. Olick & Robbins 1998). The formal cultural
memory-creating institutions in society have developed over time,
e.g. libraries, archives, monuments, and museums. In nature conser-
vation, we can add monitoring programmes as well as ensuing policy
interpretations leading to national, regional, and local conservation

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
0
3
1

130 Hilding-Rydevik et al.

programmes. National parks and nature reserves may be viewed as


monuments of the past. Ecology and nature conservation research
and researchers are of course also important parts of mnemonic
practices. Whatever the archive, there are protocols for testing its
value relative to other information, and to evaluate it against other
aspects and contexts. This is the fundamental task of historians (viz.
Bloch 2015).
Another complex example of preserving essential memories for
the future deals with the disposal of spent nuclear fuel (IAEA 1999,
2001). This issue is included in the process of planning and design-
ing the final repositories for used radioactive nuclear fuel in both
Sweden and Finland. The waste needs to be stored for more than
100,000 years in order to prevent harmful effects on humans and
the environment. The aim is to design safe storage facilities block-
ing human intrusion since the material could potentially be used
for constructing nuclear bombs. The issue is, however, complicated
since safe construction to prevent human intrusion is in opposition
to the possible need of future generations to use the spent nuclear
fuel as a resource (SOU 2007). Copper canisters with spent fuel will
be stored at 500 meters’ depth in bedrock. Once sealed, any traces
above ground must be difficult to spot. How can social memory pro-
cesses and institutions be created to ensure that future societies can
recognise why these repositories were built, where they are situated,
and what they contain in terms of substances and technology? Which
signs, languages, and media can transfer such memories in an under-
standable and meaningful way far into the future? The solutions need
to consider technological and broader social and societal aspects.
Areas of research needed to explore this issue include archaeology,
architecture, archival research, philosophy, history, art, science, and
semiotics (SOU 2015).
Nora (1989) discusses how ‘sites of memory’ (lieux de mem-
oire) are important for shaping memories. Visualising and display-
ing the past can be done through physical objects, such as buildings
and statues, or through storytelling and ceremonies. Nuttall (1992)

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
1
3

Baselines and the Shifting Baseline Syndrome 131

introduces the concept of ‘memoryscapes’ to emphasise the rela-


tion between memory and landscapes, where places become mental
images of the environment. Memories change as the world around
us changes. This makes any memoryscape a dynamic phenomenon,
resulting from individual experiences as well as from cultural and
political processes (Ullberg 2013: 178) and interactions with the
environment.

5.3.4 Further Implications of Memory Research and SBS


Memory research highlights that memories, individual and col-
lective, are constructed through complex interactions of individual
and relational processes (humans-humans, humans-surroundings).
This does not imply that accounts of past conditions are necessarily
untrue or invented; however, they are subject to change, relational,
and sometimes contested. Memories tap into the broader historical
narratives of specific societies (and beyond) which provide contexts
and vocabulary. They are also reflections of how the present under-
stands the past (Figure 5.2). For instance, when asking an elderly fish-
erman about the sizes of the catches in ‘the olden days’, chances are
that the answer would suggest that catches were larger and the fish
more plentiful compared with today. Such an account would prob-
ably not be false, but a product of many different societal processes,
not least relating to the current general discussion on biodiversity
loss and climate change (Figure 5.3). Memory is not an exact and
concrete record. It is part of a memoryscape where official ‘truths’,
societal developments, and present-day perceptions work as a com-
plex web through which what is remembered is filtered and given
meaning, on both individual and societal levels.
The SBS suggests that there is a shift in our perceptions of the
state of nature, relative to a baseline, and that this constitutes a ‘syn-
drome’. A syndrome is defined in the Merriam-Webster online dic-
tionary as: ‘a group of signs and symptoms that occur together and
characterize a particular abnormality or condition; or a set of concur-
rent things (as emotions or actions) that usually form an identifiable

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
2
3
1

132 Hilding-Rydevik et al.

Figure 5.2 Ideas and memories are shaped through many different
sources, e.g., art and paintings portraying a pristine nature. A black-and-
white version of this figure will appear in some formats. For the colour
version, please refer to the plate section.
Artwork by Jan Brueghel the Younger – 1601–78. Public domain via
Wikimedia Commons.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
3
1

Baselines and the Shifting Baseline Syndrome 133

Figure 5.3 Memories of extreme catches can influence the perception


of trends or earlier states.
Artwork by Pieter Brueghel the Elder – 1526/1530–69. Public domain
via Wikimedia Commons.

pattern’. Should the fact that individuals, groups, and societies lose
track of the ever-changing state of species and ecosystems be viewed
as an abnormality? Memory research demonstrates that a character-
istic of individuals and societies is that they remember some things
and forget others. This suggests that derogatory labels such as amne-
sia, blindness, and illusion are not appropriate for memory loss, and
questions if there actually is a ‘syndrome’ in SBS. More attention has
to be devoted to the act of remembering, as well as to when and if the
mix of remembering and forgetting does become a societal and col-
lective syndrome and abnormality.
SBS research has highlighted the importance of memory
research to the human-nature relationship. In our context, the SBS
research raises normative questions regarding what it is in the envi-
ronment that different societies need to remember and which mne-
monic practices are required to do so. What reasons are there for

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
4
3
1

134 Hilding-Rydevik et al.

remembering, for whom, for how long, and with what efforts? In the
quest for sustainable societies, these questions need further atten-
tion, both in research and in policy.

5.4 Conclusions
This chapter stresses the need for historical and multidisciplinary
awareness when setting goals and baselines in conservation, and for
understanding the complex interaction between humans and other
bio-geo-chemical aspects of the world. Such awareness is a starting
point for overcoming the nature-culture divide that is often unex-
amined in conservation research and practice. In multidisciplinary
collaboration, concepts are important vehicles, and it is through an
ongoing and critical reflection of concepts and disciplinary prac-
tices that new interdisciplinary contexts and integrative knowledge
emerge. Concepts, however, also shape and reproduce our under-
standing of the world around us, and produce normative values and
standards of their own. For instance, the concept of a ‘pristine nature’
baseline may appeal so much to common sense that it prevents cul-
tural and historical reflection on human impact and naturalness (De
Vries 2005). Critically scrutinising frames of reference and discipli-
nary concepts as creating opportunities to communicate across dis-
ciplines is important in furthering nature conservation beyond the
nature-culture divide (Setten, Stenseke, & Moen 2012). In this book,
several historical-ecological concepts are discussed, aimed at inte-
grating different concepts and disciplines, for example those of bio-
cultural diversity (Chapter 7) and memoryscape (this chapter).
This chapter discusses the need for discerning and interpreting
human impact, both beyond and within what can be remembered.
Remembering and forgetting are essential and ongoing processes,
both individually and societally, and they shape and influence the
way we create meaning, identity, and our views of the present.
Understanding how individual and collective memories are shaped
by history is also important when setting baselines for nature con-
servation. Larger numbers of people move to cities and profound

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
5
3
1

Baselines and the Shifting Baseline Syndrome 135

experiences of or relations to a diverse and rich biocultural country-


side are lost, what Atran, Medin, & Noss (2004) call the ‘extinction of
experience’. This makes remembering even more essential for policy-
and decision-makers, conservation and environmental professionals,
and the general public.
By focussing on the loosely related concepts of baselines and
the shifting baseline syndrome, and their use in nature conservation,
we have discussed how historical, socio-cultural, and political agen-
das inevitably influence nature conservation approaches and prac-
tices. Such influences are neither unexpected nor inappropriate and
apply to all societal actions, policies, or agendas. They are basically
products of the norms and values of our time and inescapably con-
structed to ‘make sense’ of the world around us. More awareness of
our frames of reference and the social construction of current nature
conservation strategies would give a better contextual understand-
ing of conservation practice and environmental politics. This would
support better-informed decisions, procedures, and activities in con-
servation. We also argue that research conducted within the field of
historical ecology would offer more nuanced and multidisciplinary
frames of reference.

References
Al-Abdulrazzak, D., Naidoo, R., Palomares, M. L. D., & Pauly, D. (2012). Gaining
perspective on what we’ve lost: the reliability of encoded anecdotes in histori-
cal ecology. PLoS ONE 7(8), e43386.
Atran, S., Medin, D., & Noss, N. (2004). Evolution and devolution of knowledge: a
tale of two biologies. Journal of the Royal Anthropological Institute, 10(2),
395–420.
Balint, P. J., Stewart, R. E., Desai, A., & Walters, L. C. (2011). Wicked Environmental
Problems. Managing Uncertainty and Conflict. Washington, DC: Island Press.
Barthel, S., Folke, C., & Colding, J. (2010). Social-ecological memory in urban gar-
dens – retaining the capacity for management of ecosystem services. Global
Environmental Change, 20 (2010), 255–65.
Belvedresi, R. E. (2014). Book review symposium. Review of the book The
Collective Memory Reader. Memory Studies, 7(1), 108–31.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
6
3
1

136 Hilding-Rydevik et al.

Berger, P. & Luckmann, T. (1966). The Social Construction of Reality. New York:
Penguin Books.
Bishop, K. H., Beven, K., Destouni, G., Abrahamsson, K., Andersson, L., Johnson,
R. K., Rohde, J., & Hjerdt, N. (2009). Nature as the ‘natural’ goal for water
management: a conversation. Ambio, 38(4), 209–14.
Bloch, M. (2015). The Historian’s Craft. Manchester: Manchester University Press.
Originally published (1949) as Apologie pour l’Histoire ou Métier d’Historien.
Paris: Armand Colin.
Brinck, E. & Frost, C. (2009). Evaluation of amendments used to prevent sodifica-
tion of irrigated fields. Applied Geochemistry, 24(11), 2113–22.
Brown, C. J. & Trebilco, R. (2014). Unintended cultivation, shifting baselines,
and conflict between objectives for fisheries and conservation. Conservation
Biology, 28(3), 677–88.
Bull, J. W., Gordon, A., Law, E. A., Suttle, K. B., & Milner-Gulland, E. J. (2014).
Importance of baseline specification in evaluating conservation interven-
tions and achieving no net loss of biodiversity. Conservation Biology, 28(3),
799–809.
Bull, J. W., Gordon, A., Watson, J. E. M., & Maron, M. (2016). Seeking convergence
on the key concepts in ‘no net loss’ policy. Journal of Applied Ecology, 53(6),
1686–93.
Burr, V. (1995). An Introduction to Social Constructionism. London: Routledge.
Campbell, L. M., Gray, N. J., Hazen, E. L., & Shackeroff, J. M. (2009). Beyond
baselines: rethinking priorities for ocean conservation. Ecology and Society,
14(1), 14.
Cardinale, B. J., Duffy, J. E., Gonzalez, A., Hooper, D. U., Perrings, C., Venail, P.,
Narwani, A., Mace, G. M., Tilman, D., Wardle, D. A., Kinzig, A. P., Daily, G.
C., Loreau, M., Grace, J. B., Larigauderie, A., Srivastava, D. S., & Naeem, S.
(2012). Biodiversity loss and its impact on humanity. Nature, 486, 59–67.
CBD. (2007). CBD Guidelines on Biodiversity and Tourism Development.
Secretariate of the Convention on Biological Diversity, February (2007).
CBD, UNEP.
Clavero, M. (2014). Shifting baselines and the conservation of non-native species.
Conservation Biology, 28(5), 1434–6.
Cook, B. & Kothari, U. (2001). Participation: The New Tyranny? London: Zed Books.
Comer, P. J. (1997). Letter. Conservation Biology, 11(2), 301–3.
Corlett, R. T. (2015). The Anthropocene concept in ecology and conservation.
Trends in Ecology and Evolution, 30(1), 36–41.
Costa, P. M., Stuart, M., Pinard, M., & Phillips, G. (2000). Elements of a certi-
fication system for forestry-based carbon offset projects. Mitigation and
Adaptation Strategies for Global Change, 5(1), 39–50.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
7
3
1

Baselines and the Shifting Baseline Syndrome 137

Dallimer, M., Tinch, D., Acs, S., Hanley, N., Southall, H. R., Gaston, K. J., &
Armsworth, P. R. (2009). 100 years of change: examining agricultural trends,
habitat change and stakeholder perceptions through the 20th century. Journal
of Applied Ecology, 46(2), 334–43.
Daw, T. M. (2010). Shifting baselines and memory illusions: what should we
worry about when inferring trends from resource user interviews? Animal
Conservation, 13(6), 534–5.
Denevan, W. M. (1992). The pristine myth: the landscape of the Americas in 1492.
Annals of the Association of American Geographers, 82(3), 369–85.
De Vries, D. (2005). Choosing your baseline carefully: integrating historical and
political ecology in the evaluation of environmental intervention projects.
Journal of Ecological Anthropology, 9(1), 35–50.
Downs, P., Singer, M., Orr, B., Diggory, Z., & Church, T. (2011). Restoring ecologi-
cal integrity in highly regulated rivers: the role of baseline data and analytical
references. Environmental Management, 48(4), 847–64.
Drew, J., Philipp, C., & Westneat, M. W. (2013). Shark tooth weapons from the
19th century reflect shifting baselines in the Central Pacific predator assem-
blies. PLoS ONE, 8(4), e59855.
EU Directive (2000). Directive 2000/60/EC of the European Parliament and of the
Council of 23 October 2000 Establishing a Framework for Community Action
in the Field of Water Policy. Official Journal of the European Communities.
L 327, 1–73.
Ellis, E. C. (2011). Anthropogenic transformation of the terrestrial biosphere.
Philosophical Transactions of the Royal Society A, 369(1938), 1010–35.
Escobar, A. (1999). After nature: steps to an antiessentialist political ecology.
Current Anthropology, 40(1), 1–30.
Ferraro, P. J. & Pattanayak, S. K. (2006). Money for nothing? A call for empir-
ical evaluation of biodiversity conservation investments. PLoS Biology,
4(4), 482–8.
Foster, N. L., Foggo, A., & Howell, K. L. (2013). Using species-area relationships
to inform baseline conservation targets for the deep north east Atlantic. PLoS
ONE, 8(3), e58941.
Gavriely-Nuri, D. (2014). Collective memory as a metaphor: the case of speeches
by Israeli prime ministers 2001–2009. Memory Studies, 7(1), 46–60.
Gergen, K. J. (2001), Social Construction in Context. Sage Publications.
Golding, L. A., Timperley, M. H., & Evans, C. W. (1997). Non-lethal responses
of the freshwater snail Potamopyrgus antipodarum to dissolved arsenic.
Environmental Monitoring and Assessment, 47(3), 239–54.
Gómez-Pompa, A. & Kaus, A. (1992). Taming the wilderness myth. BioScience,
42(4), 271–379.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
8
3
1

138 Hilding-Rydevik et al.

Gray, T. N. E., Phan, C., Pin, C., & Prum, S. (2012). Establishing a monitoring
baseline for threatened large ungulates in eastern Cambodia. Wildlife Biology,
18(4), 406–13.
Haila, Y. (1997). A ‘natural’ benchmark for ecosystem function. Conservation
Biology, 11(2), 300–1.
Haila, Y. (2000). Beyond the nature-culture dualism. Biology and Philosophy,
15(2), 155–75.
Halbwachs, M. (1992). On Collective Memory. Transl./ed. L. A. Coser. Chicago:
University of Chicago Press.
Healy, C. & Tumarkin, M. (2011). Social memory and historical justice.
Introduction Memory Studies, 4(1), 3–12.
Hoffmann, B. D., Auina, S., & Stanley, M. C. (2014). Targeted research to improve
species management: Yellow Crazy ant Anoplolepis gracilipes in Samoa. PLoS
ONE, 9(4), e95301.
Hooper, D. U., Carol Adair, E., Cardinale, B. J., Byrnes, J. E. K., Hungate, B. A.,
Matulich, K. L., Gonzalez, A., Duffy, J. E., Gamfeldt, L., & O’Connor, M. I.
(2012). A global synthesis reveals biodiversity loss as a major driver of ecosys-
tem change. Nature, 486, 105–9.
Hornborg, A. & Pálsson, G., eds. (2000). Negotiating Nature: Culture, Power, and
Environmental Argument. Lund, Sweden: Lund Studies in Human Ecology,
Lund University Press.
Hunter, M. (1996). Benchmarks for managing ecosystems: are human activities
natural? Conservation Biology, 10(3), 695–7.
IAEA. (1999). Maintenance of Records for Radioactive Waste Disposal. Report
IAEA-TECDOC-1097. International Atomic Energy Agency.
IAEA. (2001). Waste Inventory Record Keeping Systems (WIRKS) for the
Management and Disposal of Radioactive Waste. Report IAEA-TECDOC-
1222. International Atomic Energy Agency.
Jackson, J. B. C. & Alexander, K. E. (2011). Introduction: the importance of shifting
baselines. In J. B. C. Jackson, K. E. Alexander, & E. Sala, eds., Shifting Baselines.
The Past and the Future of Ocean Fisheries. Washington, DC: Island Press,
pp. 1–7.
Jackson, J. B. C., Alexander, K. E., & Sala, E. eds. (2011). Shifting Baselines. The
Past and the Future of Ocean Fisheries. Washington, DC: Island Press.
Josefsson, T. (2009). Pristine Forest Landscapes as Ecological References. Human
Land Use and Ecosystem Change in Boreal Fennoscandia. Uppsala: Swedish
University of Agricultural Sciences.
Josefsson, T., Bergman, I., & Östlund, L. (2010). Quantifying Sami settlement and
movement patterns in northern Sweden 1700–1900. Arctic, 63(2), 141–54.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
9
3
1

Baselines and the Shifting Baseline Syndrome 139

Josefsson, T., Hörnberg, G., & Östlund, L. (2009). Long-term human impact and
vegetation changes in a boreal forest reserve: implications for the use of pro-
tected areas as ecological references. Ecosystems, 12(6), 1017–36.
Kai, Z., Woan, T. S., Jie, L., Goodale, E., Kitajima, K., Bagchi, R., & Harrison, R.
D. (2014). Shifting baselines on a tropical forest frontier: extirpations drive
declines in local ecological knowledge. PLoS ONE, 9(1), e86598.
Kenny, M. G. (1999). A place for memory: the interface between individual
and collective history. Comparative Studies in Society and History, 41(3),
420–37.
Knowlton, N. & Jackson, J. B. C. (2008). Shifting baselines, local impacts, and
global change on coral reefs. PLoS Biology, 6(2), e54.
Latour, B. (2004). Politics of Nature. How to Bring the Sciences into Democracy.
Cambridge, MA: Harvard University Press.
Lélé, S. & Norgaard, R. B. (1996). Sustainability and the scientist’s burden.
Conservation Biology, 10(2), 354–65.
Lewis, S. L. & Maslin, M. A. (2015). Defining the Anthropocene. Nature, 519,
171–80.
Linde, C. (1997). Narrative: experience, memory, folklore. Journal of Narrative
and Life History, 7(1), 281–90.
Linde, C. (2000). The acquisition of a speaker by a story: how history becomes
memory and identity. Ethnos, 28(4), 608–32.
Lorimer, J., Sandom, C., Jepson, P., Doughty, C., Barua, M., & Kirby, K. J. (2015).
Rewilding: science, practice, and politics. Annual Review of Environment and
Resources, 40(1), 39–62.
Maron, M., Bull, J. W., Evans, M. C., & Gordon, A. (2015). Locking in loss: baselines
of decline in Australian biodiversity offset policies. Biological Conservation,
192, 504–12.
Maron, M., Hobbs, R. J., Moilanen, A., Matthews, J. W., Christie, K., Gardner,
T. A., Keith, D. A., Lindenmayer, D. B., & McAlpine, C. A. (2012). Faustian
bargains? Restoration realities in the context of biodiversity offset policies.
Biological Conservation, 155 (2012), 141–8.
McClenachan, L., Feretti, F., & Baum, J.K. (2012). From archives to conserva-
tion: why historical data are needed to set baselines for marine animals and
ecosystems. Conservation Letters, 5(5), 349–59.
McDonald-Madden, E., Gordon, A., Wintle, B. A., Walker, S., Grantham, H.,
Carvalho, S., Bottrill, M., Joseph, L., Ponce, R., Stewart, R., & Possingham, H.
P. (2009). ‘True’ conservation progress. Science, 323(5910), 43–4.
Michelutti. N., McCleary, K. M., Antoniades, D., Sutherland, P., Blais, J. M.,
Douglas, M. S. V., & Smol, J. P. (2013). Using paleolimnology to track the

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
0
4
1

140 Hilding-Rydevik et al.

impacts of early Arctic peoples on freshwater ecosystems from southern


Baffin Island, Nunavut. Quaternary Science Reviews, 76, 82–95.
Natlandsmyr, B. & Hjelle, K. L. (2016). Long-term vegetation dynamics and land-
use history: providing a baseline for conservation strategies in protected Alnus
glutinosa swamp woodlands. Forest Ecology and Management, 372, 78–92.
Nora, P. (1989). Between memory and history: les lieux de memoire.
Representations, 26, 7–24.
Nuttall, M. (1992). Arctic Homeland: Kinship, Community and Development in
Northwest Greenland. London: Belhaven.
Nykvist, B. & von Heland, J. (2014). Social-ecological memory as a source of gen-
eral and specified resilience. Ecology and Society, 19(2), 47.
Olick, J. K. & Robbins, J. (1998). Social memory studies: from ‘collective mem-
ory’ to the historical sociology of mnemonic practices. Annual Review of
Sociology, 24, 105–40.
Olick, J. K., Vinitzky-Seroussi, V., & Levy, D. (2014). Response to our critics.
Memory Studies, 7(1), 131–38.
Ortmann, G. (2010). On drifting rules and standards. Scandinavian Journal of
Management, 26(2), 204–14.
Papworth, S. K., Rist, J., Coad, L., & Milner-Gulland, E. J. (2009). Evidence for shift-
ing baseline syndrome in conservation. Conservation Letters, 2(2), 93–100.
Pauly, D. (1995). Anecdotes and the shifting baseline syndrome of fisheries. TREE,
10(10), 430.
Rautio, A.-M., Josefsson, T., & Östlund, L. (2014). Sami resource utilization and
site selection: historical harvesting of inner bark in northern Sweden. Human
Ecology, 42(1), 137–46.
Reed, M. S. (2008). Stakeholder participation for environmental management: a
literature review. Biological Conservation, 141(10), 2417–31.
Rice, J. (2013). Further beyond the Durkheimian problematic: environmental
sociology and the co-construction of the social and the natural. Sociological
Forum, 28(2), 236–60.
Roediger, H. L., Marsh, E. J., & Lee, S. C. (2002) Varieties of memory. In D. L Medin
& H. Pashler, eds., Stevens Handbook of Experimental Psychology: Memory
and Cognitive Processes, vol. 2, 3rd ed. New York: John Wiley and Sons,
pp. 1–41.
Rothberg, M. (2009). Multidirectional Memory: Remembering the Holocaust in
the Age of Decolonisation. Stanford, CA: Stanford University Press.
Ryan, J. C. (2013). Botanical memory: exploring emotional recollections of native
flora in the southwest of western Australia. Emotion, Space and Society,
8, 27–38.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
1
4

Baselines and the Shifting Baseline Syndrome 141

Sáenz-Arroyo, A., Robert, C. M., Torre, J., Carino-Olvera, M., & Enriquez-Andrade,
R. R. (2005). Rapidly shifting environmental baselines among fishers of the
Gulf of California. Proceedings of the Royal Society B, 272(1575), 1957–62.
Setten, G. & Austrheim, G. (2012). Changes in land use and landscape dynamics in
mountains of northern Europe: challenges for science, management and con-
servation. International Journal of Biodiversity Science, Ecosystem Services
and Management, 8(4), 287–91.
Setten, G., Stenseke, M., & Moen, J. (2012). Ecosystem services and landscape man-
agement: three challenges and one plea. International Journal of Biodiversity
Science, Ecosystem Services and Management, 8(4), 305–12.
Sörlin, S. (1991). Humanekologi: vägar till överlevnad. Umeå: Forum för tvärvet-
enskap, Umeå universitet.
SOU 2007: 38. Kunskapsläget på kärnavfallsområdet 2007. Nu levandes ansvar,
framtida generationers frihet. Stockholm: Statens råd för kärnavfallsfrågor
SOU 2015: 11. Kunskapsläget på kärnavfallsområdet 2015. Kontroll, dokumenta-
tion och finansiering för ökad säkerhet. Stockholm: Kärnavfallsrådet.
Terry, J. (2013). ‘When the sea of living memory has receded’: cultural memory
and literary narratives of the Middle Passage. Memory Studies, 6(4), 474–88.
Ullberg, S. (2013). La Inundación – katastrofer och minnets politik i Argentina. In
M. Viktorin & C. Widmarrd, eds., Antropologi och tid. Stockholm: Svenska
sällskapet för Antropologi och Geografi, pp. 175–92.
Valinia, S., Hansen, H.-P., Futter, M. N., Sriskandarajah, N., & Fölster, J. (2012).
Problems with the reconciliation of good ecological status and public partici-
pation in the Water Framework Directive. Science of the Total Environment,
433(2012), 482–90.
Wilson, R. A. (2005). Collective memory, group minds, and the extended mind
thesis. Cognitive Processing, 6, 227–36.
Zelizer, B. (1995). Reading the past against the grain: the shape of memory studies.
Critical Studies in Mass Communication, 12(2), 2014–239.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
2
4
1

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:35, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.005
3
4
1

Part II Approaches:
Concepts and Methods

Downloaded from https://www.cambridge.org/core. The University of British Columbia Library, on 02 Aug 2018 at 17:43:15, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
4
1

Downloaded from https://www.cambridge.org/core. The University of British Columbia Library, on 02 Aug 2018 at 17:43:15, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
5
4
1

6 Concepts for Integrated


Research in Historical Ecology
Ove Eriksson, Anneli Ekblom, Paul Lane,
Tommy Lennartsson, and Karl-Johan Lindholm

6.1 Introduction
The aim of historical ecology, as a research programme, is to be
inherently integrative, crossing over boundaries that divide what are
usually considered as separate academic disciplines: ecology and evo-
lutionary biology from the natural sciences, and mainly archaeology,
anthropology, human geography, and history from the social sciences
and humanities (Crumley 2007; Meyer & Crumley 2011). In this
chapter, we seek new ways of improving current frameworks for dis-
ciplines engaged in integrated research, introduced through historical
ecology (Crumley 1994, 2007; Balée 2006; Meyer & Crumley 2011;
Szabó & Hedl 2011) and other similar research areas. Our overall aim
is a reappraisal of influential concepts, applied in previous research
in landscape ecology and archaeology, such as adaptation and niche
construction. These concepts have been borrowed from evolutionary
biology and ecology for the analysis and understanding of the socio-
environmental interface, and have had a strong influence on research
directions in environmental anthropology and archaeology (see, e.g.
the articles in Kendal, Tehrani, & Odling-Smee 2011). Hence, we
seek to explore concepts that are already points of reference among
the different fields. Our main argument is that it is possible to refor-
mulate concepts, such as niche construction and adaptation, to
create a new framework that draws upon concepts mainly derived
from social history, for example entanglement, assemblage, practice,
and structuration. An integration of these different concepts will
not only strengthen historical ecology, but also add to our current

145

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
6
4
1

146 Eriksson et al.

understanding of the history of interactions among human and other


biological agents active in a landscape.

6.2 Background
In the 1970s and 1980s, ideas from the natural sciences were rather
rigidly applied in Anglophone European archaeology and in some
areas of North American anthropology (e.g. Rappaport 1968; Vayda
1969; Dunnell 1971; Watson, LeBlanc, & Redman 1971; Renfrew
1972; Renfrew, Rowlands, & Segraves 1982). This eventually
prompted a reaction against the borrowing of concepts and concep-
tual frameworks from the natural sciences (Ellen 1982; Hodder 1982;
Wylie 1985; Marcus & Fischer 1986; Shanks & Tilley 1987; Escobar
1996; see Davies 2013 for further discussion). The scepticism was at
least partly derived from an awareness of agency and historical con-
tingency, neither of which were accommodated within the ecological
models of human behaviour and society that then dominated archae-
ology and anthropology. Practice theory, as articulated by Pierre
Bourdieu (1977), especially his concept of ‘habitus’, and Anthony
Giddens’ (1984) theory of structuration, proposes that both change
and stability are inherent to relations among individual agents.
This concept was particularly influential in stimulating early post-
processual critiques in British archaeology (e.g. Hodder 1982; Shanks &
Tilley 1987). In North America, initial critique coincident with these
trends entailed recognition of the importance of historical contin-
gency in social processes, and the imbalances in power and access
to resources that it can engender, as inspired by anthropologists (e.g.
Wolf 1982; Sahlins 1987), and archaeologists (e.g. Gero & Conkey
1991; Spector 1993). One strand of critique drew on Hegelian ideas
about the dialectical nature of the relationship between humans and
their environments; this trans-temporal approach was particularly
influential in shaping early historical ecology (e.g. Crumley 1987;
Marquardt 1992). Studies inspired by Marx and Engels also offered
alternative perspectives on human-environment relations (e.g. Fried
1967; Friedman & Rowlands 1978; Gilman 1981; Kirch 1984; Spriggs

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
7
4
1

Concepts for Integrated Research in Historical Ecology 147

1984) by raising awareness of their political dimensions and the


variable power of individual actors and institutions to effect change
according to their structural position within society. The differen-
tial distribution of power within society, and its consequences both
for individuals and groups in shaping their social worlds and the
physical landscape remains important (e.g. Robbins 2012; Earle &
Spriggs 2015).
Meanwhile, studies of landscape ecology largely ignored the
influence of humans (see critical review in Szabó 2010), in the shap-
ing of ecological relations (e.g. Widgren 2012). These studies have
generally displayed a poor understanding of the importance of his-
torical relations among humans, specifically as creative agents – not
merely as destructive ones (Kricher 2009) – and other biological agents
in shaping landscapes, resulting in ideas of landscapes as pristine or
degraded (see critical discussion in Denevan 2011). Consequently,
many landscape studies have focused either on interactions between
biological agents and the physical landscape, with little interest in
the interaction with human social relations, or on human social rela-
tions as expressed and experienced in a relatively static physical land-
scape, where other biological agents have been assigned little or no
impact on these relations. There are, of course, some notable excep-
tions. In archaeology, for instance, Karl Butzer’s synthesis (1982) and
regionally specific studies (e.g. Butzer 1996, 2005) are examples of
landscape-oriented research aimed at bridging this divide. The study
of the historical ecology of the landscapes of southern Scandinavia
(Berglund 1991) is another example where aspects of both social
history and landscape ecology have been integrated. Even in these,
however, concepts of agency, actor networks, the materiality of envi-
ronment, and the constitution of an environmental habitus are either
absent or marginally developed.
Studies in historical ecology should and must integrate both
historical and ecological phenomena. As shown throughout this book,
the integration of landscape ecology, archaeological long-term per-
spectives, and social theory is crucial for understanding sustainable

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
8
4
1

148 Eriksson et al.

resource management and the long-term effects of management deci-


sions and strategies. Accordingly, an overall key concept in this chap-
ter is that landscapes are constituted by relations among biological
and non-biological agents and processes, and are shaped and reshaped
by the experience, knowledge, and practice of individuals and social
groups. Our position is that landscapes are more than spatial mani-
festations of these human-environment relations, they are also active
agents in the constitution of these relations and their reproduction
and transformation over time. This discourse is, however, still trou-
bled by a legacy of separating the natural sciences and the human-
ities that goes back to the late nineteenth century (Foucault 1970;
Latour 1999).

6.3 Changing Concepts


Recent decades have witnessed an increasing convergence of efforts
across the disciplines aimed at dispensing with conventional dichoto-
mies, such as ‘nature’ and ‘culture’, ‘humans’ and ‘the environment’.
Aspects of this trend can be attributed to renewed scholarly interest
in the relations between physical landscapes and the biological world
(including humans) as agents, as well as animal-centred perspectives
on such concepts as domestication, place, and identity (e.g. Anderson
1997; Wolch & Emel 1998). Other developments within the social
sciences, of pertinence here, concern the manner in which the rela-
tions between human and biological agents are now considered to
create networks of knowledge of the world (Law & Mol 2002). This
field of study originates in science and technology (e.g. Haraway 1991;
Latour 1999, 2000, 2005) via ‘hybrid geography’ (Callon & Law 1995;
Whatmore 2002; Hinchcliffe 2007) to a more recent fusion of geog-
raphy and political theory (Thrift 2008; Bennet 2010). The perspective
allows for agency, not just in humans, but also in the biological and
physical worlds. These ideas also find inspiration in earlier philosoph-
ical proposals by Deleuze and Guattari (2010) that introduced the con-
cept of knowledge as rhizome-like (though here we use the concept of
network). In the study of a network of relations, one has to dissolve

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
9
4
1

Concepts for Integrated Research in Historical Ecology 149

the predefinitions and demarcations entailed in the separation of sci-


ences into categories such as ‘nature’ and ‘culture’. Both historical
ecology and landscape studies carry this potential precisely because
they can incorporate many different perspectives and meanings, and
actively seek to bridge the artificial divide between what is cultural
and what is natural. Hence, the landscape provides possibilities to
integrate separate sciences, and also to accommodate diversity in
perspectives by placing more focus on what can be considered ‘ecol-
ogies of practice’, i.e. an understanding of how practices, undertaken
by individuals connected by networks such as social and political
structures and knowledge systems, shape landscapes and are in turn
shaped by these self-same landscapes (Nyerges 1997; Balée 2006).
Within archaeology, this conceptual reframing has been explored
by Hodder (2012), among others, using the concept of human-thing
‘entanglement’. In broad terms, Hodder (2012) identifies three sets of
paired relationships: between humans and things, things and things,
and humans and humans, that are simultaneously and dialectically
ones of dependence and dependency. The idea of ‘dependence’ carries
with it dual notions of a relationship (whether just between humans,
between things, or between humans and things) enabling something
else, while also making each pair contingent on the other. Regarding
‘dependency’, Hodder uses this concept to indicate relationships that
create conditions of reliance between each component of the pair,
and so constrain and foreclose future options and directions of devel-
opment. Hodder also emphasises the centrality of time and tempor-
ality to human-thing entanglement, both in terms of scheduling and
sequence, but also of the non-linear unfolding of trajectories and the
responses evoked by unpredictable conjunctures between humans
and things at multiple scales. In his terms, ‘the determinative com-
ponent in entanglements that pushes them in certain directions is
not the economic base, the ecology nor the infrastructure; but nei-
ther is it ideology, systems of meaning nor the superstructure. Rather
the entangling itself has a tautness that channels and directs humans
and things as they go about their daily business of dependence and

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
0
5
1

150 Eriksson et al.

co-dependence’ (Hodder 2012: 208, author’s emphasis). As will be


exemplified later in this chapter, this concept may have potential
also for the reframing of landscape ecology.
When environmental research began in the early twentieth cen-
tury, one dominant ecological paradigm was that of the ‘balance of
nature’, an idea with roots in early Greek philosophy (Egerton 1973;
Kricher 2009). One of the most influential early treatments of veg-
etation and landscape ecology was by Clements (1916), who argued
that with time any vegetation system would reach a climax stage,
i.e. a state where the vegetation was in balance (i.e. in equilibrium)
with the surrounding physical environment. Clements was heavily
influenced by holistic ideas and he regarded a climax community as
a ‘superorganism’, with a determined ontogeny analogous to organ-
isms. Although the idea of plant succession was not new, and ear-
lier treatments by Warming (1895) and Cowles (1899) had introduced
directed changes in plant communities, Clements’ holistic ideas on
plant succession became enormously important, particularly in the
United States (e.g. Worster 1994), and they influenced the develop-
ment of systems ecology (Odum 1983; O’Neill et al. 1986). Clements’
ideas were, however, subjected to strong criticism as early as the
1920s, mainly by Gleason (1926), who, by introducing the ‘individu-
alistic concept of ecology’, challenged the idea of landscape as con-
sisting of discrete communities behaving in a systemic manner. In
Gleason’s view, vegetation is composed of assemblages of species,
reflecting each individual species’ features and influencing their dis-
persal and capacity for maintaining populations. Later development
of landscape and vegetation ecology was strongly influenced by this
reasoning, viewing landscape systems as dynamic in space and time
(e.g. Watt 1947; Whittaker 1975), thus in a way merging both tempo-
ral (historical) and spatial dimensions of landscapes (e.g. Forman &
Godron 1986).
Most ecosystems are affected by disturbances of various
kinds (e.g. Pickett & White 1985; Hubbell & Foster 1986; Wu &
Loucks 1995), a perspective that has encouraged a revision of the

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
1
5

Concepts for Integrated Research in Historical Ecology 151

understanding of ‘mosaic landscapes’, where the combination of dis-


turbances and the temporal and spatial variability among them cre-
ates a landscape in flux. The emerging view of ecological systems as
being built up of individuals or populations of species that are con-
stantly changing without ever reaching a state of balance, is similar
to the revision of the understanding of entities such as ‘culture’ or
‘society’ in the human sciences. The individualistic concept in land-
scape ecology opened up a view of landscapes as inherently chan-
ging and historical, without assuming an overall goal or direction
of change. This conceptual framework also opens up the possibility
for reintroducing humans into landscape ecology as creative agents,
shaping conditions and fostering heterogeneity.

6.4 A New Framework for Understanding


Socio- environmental Relations
The challenge for historical ecology (as for human geography, envi-
ronmental history, and landscape archaeology) is to explore and rep-
resent landscapes created through interaction among humans, the
biological world, and physical processes. This remains a difficult task
as it requires a conceptual reframing, not only of the study of such
interactions, but also of how we create knowledge about them. We
will use the two familiar concepts of niche construction and adap-
tation to exemplify how, if re-conceptualised within the frame of
entanglement, these can represent and explain socio-environmental
relations as a continuum. We will also illustrate this continuum
through the example of landscape domestication.

6.4.1 Revisiting Niche Construction


The niche concept has a long tradition in ecology (for an overview,
see e.g. Chase & Leibold 2003), and there is a plethora of literature
discussing the definition and measurement of ecological niches (e.g.
Soberón 2007; Godsoe 2010). The past few years have also seen a
reinvigorated interest in the concept on the part of archaeologists (see
reviews in Broughton, Cannon, & Bartelink 2010; Laland & O’Brien

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
2
5
1

152 Eriksson et al.

2011; Dieter 2012). Niche construction theory, as well as other


ideas, has been used to explain the emergence of agriculture (Smith
2011a, 2011b; Laland & O’Brien 2012; O’Brien & Laland 2012), and
has often influenced discussions of the processes of domestication
more broadly. For the purpose of our discussion, we refer to the basic
definitions proposed by Hutchinson (1957), distinguishing between
the fundamental and the realised niche. The fundamental niche rep-
resents a set of environmental conditions which permits a species
to persist indefinitely, whereas a realised niche is a subset of these
conditions which is actualised by a particular species at a specific
time and place. Interaction among species, e.g. the competition for
resources, or general environmental changes, may induce niche
shifts, defined as changes in the realised niche, and may also initi-
ate an evolutionary process that gradually changes the fundamental
niche of a species (e.g. Pearman et al. 2008; Eriksson 2013).
Elaborated by Odling-Smee and colleagues (2003, 2013), niche
construction is defined as the process whereby organisms, through
their metabolism, activities, and behaviour, modify their own and/
or other species’ niches. Conceptually, niche construction com-
plements the idea of agents in the landscape and of entanglement
(Hodder 2012). In this respect, it is perhaps of no surprise that Charles
Darwin (1859: 74) used the metaphor of an entangled bank in the
penultimate paragraph of his seminal work, On the Origin of Species,
to convey the notion that species interact and influence the envi-
ronment in various ways, directly or indirectly influencing resources
and conditions for themselves and other species. Beavers, ants, domi-
nant forest trees, elephants, and humans are examples of species that
both construct their own environment and have a strong impact on
other species. What is important is that niche construction implies
the existence of a non-linear feedback process among the niche-con-
structing agent, the effects on the environment, the response of asso-
ciated species, and back to the niche-constructing agent. Over time,
these processes can lead to evolutionary changes in species through
natural selection or by knowledge transfer (in humans) through what

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
3
5
1

Concepts for Integrated Research in Historical Ecology 153

has been termed eco-evolutionary dynamics (Post & Palkovacs 2009)


or simply co-evolution (e.g. Rindos 1984; Thompson 1994). Homo
sapiens have been considered as the ‘ultimate niche constructing
species’ (Smith 2007: 188). Human impact on the world’s ecosys-
tems is ubiquitous, and more than 75 per cent of the Earth’s ice-free
terrestrial surface is so altered that some authors suggest defining
new ‘anthropogenic biomes’ (Ellis & Ramankutty 2008; Ellis et al.
2013). The niche-constructing ability of humans has increased over
time, and is closely related to the means and strategies by which
humans use resources, from stone tools, to the burning of fossil fuel
and laptops.
Human niche construction may also be understood in a wider
sense, relating to the evolutionary history of the species, espe-
cially referring to its cognitive abilities. Tooby and DeVore (1987)
introduced the concept ‘cognitive niche’ encapsulating humans’
behaviourally unique capacity to broaden resource capture by cause-
and-effect reasoning, learning, and cooperative action. This idea has
since been extended by several authors. Pinker (2010) stressed the
importance of language for exploiting the cognitive niche, while
Whiten and Erdal (2012) emphasised the significance of egalitarian-
ism for the development of culture. Boyd, Richerson, and Henrich
(2011) proposed the concept of ‘cultural niche’ and argued that the
process of knowledge transfer was a key component in developing
cultures: ‘cultural evolution operating over generations has gradually
accumulated and recombined adaptive elements, eventually creating
adaptive packages beyond the causal understanding of the individuals
who use them’ (Boyd et al. 2011: 10923). Together, these concepts
envisage a niche-constructing process operating on several levels, ini-
tially related to cognitive capacity per se, but ultimately manifested
as the means by which humans construct and use their environment.

6.4.2 Revisiting Adaptation


By repeating the basic proposition that biological agents (includ-
ing humans) construct conditions for each other and the landscapes

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
4
5
1

154 Eriksson et al.

with which they interact, we may take a fresh look at the concept of
adaptation that has been so mis-conceptualised in earlier landscape
research. In everyday language, adaptation refers to the process of
changing something in order to make it useful for a new purpose.
In the humanities, the idea of adaptation was embraced within eco-
logical functionalism, where cultural expressions and social organ-
isation were assumed to be functionally adapted to ensure the
survival of society. In biology, adaptation may refer to a physiological
process, although the most common biological meaning is the one
used within evolutionary biology. Here, it refers to features of organ-
isms that have evolved into their current functional role by means
of natural selection. Adaptation in this context has a dual meaning,
referring both to the adaptive feature and the process that leads to
the evolution of this feature. This process is ‘blind’, i.e. it has no pur-
pose or goal. Natural selection implies that adaptation evolves, as if
constructed with a purpose, a common misconception found in both
natural sciences and humanities.
In evolutionary biology generally, adaptation has been con-
sidered strictly in the context of selection for genetically heritable
features. There is, however, an increasing appreciation that features
that are not genetically inherited may be important in ecological and
evolutionary processes, provided that there are other means of infor-
mation transfer across generations (Wilson 2005: 21). Danchin (2013)
suggested that there are at least four means by which non-genetic
variation of features can be transmitted across generations, of which
one, the transfer of cultural variation, is relevant in the context of our
proposed framework. The key issue is that cultural variation (which
may include all aspects relevant for social life, including knowledge
of resource use) expressed in one generation, by means of non-genetic
inheritance, is transmitted to the next generation. The mechanisms
that determine this cross-generational transfer of cultural informa-
tion include learning, memory, and knowledge exchange among indi-
viduals, i.e. the key components of any cultural niche (Boyd et al.
2011). Such processes, even if dissimilar to evolutionary processes in

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
5
1

Concepts for Integrated Research in Historical Ecology 155

a neo-Darwinian sense (e.g. Mesoudi 2011), would be equally capable


of producing adaptation.
Adaptation is, therefore, here defined as features that have
evolved or developed by means of selection or intention for their cur-
rent roles, through genetically heritable variation or processes that
involve learning, memory, and knowledge transfer in order to change
or maintain a feature. Niche construction and domestication of land-
scapes, which is further discussed later in this chapter, are examples
of such processes that not only create altered niche opportunities for
numerous other species, but also environmental conditions perceived
as beneficial or detrimental to human society. They also initiate a
feedback process that over time may alter or induce further niche
construction, and also change the social relations among people and
among humans, plants, and animals. This makes the domesticated
landscape evolve over time as a relationship between the biological
world (of which humans form a part) and the physical environment.

6.5 Domesticated Landscapes


The most obvious implication of human niche construction for other
species is domestication, and we will therefore continue to illustrate
how hybridity, praxis, social transfer, and niche construction interact
in a domesticated landscape. We must, however, start by examining
the concept of domestication. The past few decades have witnessed
considerable discussion and debate on how to define domestication
and the processes by which it occurs. Archaeologists and other schol-
ars have long acknowledged that domestication of plant or animal
species is not instantaneous (e.g. Jarman & Wilkinson 1972; Rindos
1984), contrary to older descriptions (pace Childe 1928) of the adop-
tion of agriculture as a ‘revolution’. Instead, domestication was typ-
ically a drawn-out process involving a range of social, biophysical,
and biogenetic material and landscape phenomena (e.g. Zeuner 1963;
Harris 1996; O’Connor 1997; Russell 2002; Leach 2003; Zeder 2011).
Consequently, it is now widely held that observable morphological
and physiological differences among ‘domesticated’ species and their

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
6
5
1

156 Eriksson et al.

presumed ‘wild’ ancestor, at best, mark only a particular stage in


the process, as, for example in Zeuner’s five-stage model of animal
domestication (1963). There is also widespread acceptance that the
relationships established between humans and other species, whether
animal or plant, during the drawn-out process of domestication often
tend to be symbiotic rather than simply exploitative (e.g. Harris 1996;
O’Connor 1997). More recently, however, additional arguments con-
cerning domestication have been put forward which, although they
have not reduced the importance of using morphological and physio-
logical changes to identify domestication in archaeological contexts,
have encouraged alternative conceptualisations of the processes of
domestication, the agency of the actors, and even the state of being
‘domesticated’. Perhaps the most fundamental shift in the current
understanding of domestication concerns the move from the char-
acterisation of ‘wild’ and ‘domesticated’ as mutually exclusive to a
view where these terms instead are treated as states along a con-
tinuum (Rosman & Rubel 1989; Dobney & Larson 2006: 261).
Domesticated landscapes (e.g. Erickson 2006; Terrell & Hart
2006; Kareiva et al. 2007; Widgren 2012; Eriksson & Arnell 2017)
are places where humans have created environmental niches, not
only supporting human society, but also affecting many other spe-
cies. Erickson’s (2006, 2008, 2010) studies in Amazonia offer several
good examples of how humans, through modifications of the land-
scape, have created niche opportunities for many species, used both
by humans and other species. This was achieved through the open-
ing of light gaps, cultivating, fertilising, weeding, and altering water
conditions (see also Clement et al. 2015). Next, two further examples
of domesticated landscapes, from Africa and Europe, are discussed in
detail.

6.5.1 Domesticated Landscapes in Africa


There is emerging evidence that unintended actions of farmers and
herders in different parts of sub-Saharan Africa have created domes-
ticated landscapes, turning them into more productive and amenable

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
7
5
1

Concepts for Integrated Research in Historical Ecology 157

Figure 6.1 Aerial view of occupied and abandoned pastoralist bomas


in Amboseli (Kenya), January 2009, showing visual differences in their
higher nutrient content relative to background soils.
Photo: P. J. Lane.

niches for a wide range of species including, but not restricted to,
humans (domesticated African pastoralist landscapes: Figures 6.1–
6.3). Examples include the formation of African Dark Earths (AfDEs)
in the West African rainforest zone (Frausin et al. 2014; Fraser,
Frausin, & Jarvis 2015), the relationships between human settle-
ment practices and baobab tree (Adansonia digitata L.) recruitment
in Sahelian West Africa (Duvall 2007), as well as West African park-
land agro-forestry more generally (Blench 2007; Maranz 2009; cf.
Mather 2003, for a discussion of additional cultural dimensions of
West African landscape domestication). Here, the focus is placed on
the domestication of semi-arid and arid savanna landscapes in East
Africa. These commonly contain open areas (typically 0.5–1 ha. in
size) of cropped grass within a wider mosaic of woody and/or bushy
vegetation. Often known as ‘glades’, these open patches have diverse
origins and complex histories. Various natural processes including
fires, long-term droughts, ungulate density and grazing regimes, and

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
8
5
1

158 Eriksson et al.

Figure 6.2 Dung-enriched soil in a recently abandoned pastoralist


boma, Amboseli (Kenya), January 2009.
Photo: P. J. Lane.

the actions of large keystone species, such as elephants and rhinos,


can all contribute to the creation of glades and the maintenance of
grazing lawns (Archibald 2008). Once established, a series of feed-
back mechanisms comes into play that helps maintain glades by
restricting tree recruitment. Researchers from a number of discip-
lines have sought to understand these processes and their shifting
dynamics, both from an ecological perspective and in relation to the
implications for environmental management and wildlife conserva-
tion (Reid & Ellis 1995; Young, Partridge, & Macrae 1995; Augustine
2003; Augustine & McNaughton 2004). This ecological research has
confirmed that, as a result of a series of relationships of ecological

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
9
5
1

Concepts for Integrated Research in Historical Ecology 159

Figure 6.3 Managed trees (Acacia erioloba) at a pastoral well site


(Ozombu Zo Vindimba) in the Kalahari, eastern Namibia. The tree
has new and several older marks indicating that branches have been
cut. Reasons for cutting branches are to harvest browse, firewood, or
timber from trees without killing them, aiming for the full regrowth of
the tree. The local herders explain the trees’ curved shape as a result of
when the trees were younger, they were often bent down towards the
ground for feeding calves with the highly nutritious seed pods.
Photo: K. J. Lindholm.

mutualism, in some landscapes glades often mark the location of


abandoned pastoralist settlements where livestock were penned
overnight.
These relationships can be summarised as follows: the penning
of livestock results in a dense accumulation of dung, in some cir-
cumstances producing mounds up to two metres above the ground.
With the regular addition of animal urine, these deposits result in
heightened concentrations of several soil minerals, especially nitro-
gen (N) and phosphorous (P), but also magnesium (Mg), calcium
(Ca), carbon (C), and potassium (K), which are all beneficial to plant

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
0
6
1

160 Eriksson et al.

productivity. In the low-nutrient soils of southern Turkana (Kenya),


for example, soils in abandoned pastoralist settlements typically con-
tain ‘nine times more C, three times more N, and six times more P’
than found at background levels, and even twenty years after aban-
donment, nutrient levels are at least twice as high (Reid & Ellis
1995: 984). Decomposition of the accumulated dung also enhances
the water retention capacity of the soil, possibly by reducing evapo-
transpiration rates (Reid & Ellis 1995; Muchiru, Western, & Reid
2008) and rainwater penetration. Sustained and repeated occupa-
tion also leads to progressive removal of trees and bushes, which are
used for building materials and fuel, and the intense grazing of the
surrounding sward.
After abandonment, the high phosphorous and nitrogen con-
tents of the soil above former livestock pens, including both the cen-
trally located main pens and more peripheral enclosures for calves
and small stock (i.e. sheep and goats), restrict the sequence of plant
re-colonisation to certain species that can tolerate these levels. The
exact composition of these pioneer communities varies according
to different bioclimatic conditions, but in all documented cases, the
grasses are rich in nutrients and therefore attract wild and domes-
tic ungulates, whose repeated grazing helps to maintain these areas
as open, closely cropped ‘lawns’ or ‘glades’. Important forage plants
occur throughout the succession, in early stages these include
the stargrass Cynodon plectostachyus and in the later ones, for
example Pennisetum stramineum (Augustine 2003; Muchiru et al.
2008). Over time, woody biomass increases, with certain species,
such as the Umbrella thorn, Acacia tortilis, becoming particularly
dominant.
In southern Turkana, an interesting relationship of ecological
mutualism between this species and domestic animal dung has
developed. More precisely, A. tortilis seedpods are rich in protein and
are fed to young and milking livestock in the dry season when other
forage sources are scarce or of poor quality (Reid & Ellis 1995: 979),
and a system of usufruct rights governing access to different

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
1
6

Concepts for Integrated Research in Historical Ecology 161

concentrations of A. tortilis has evolved (Barrow 1990). Ingestion


of the seed stimulates germination and destroys seed-predating
bruchid beetles, while the dung layers in which they are deposited
provide a nutrient-rich seed bank (Lamprey, Halevy, & Makacha
1974). Although variations in rainfall and herbivory can further
influence seedling establishment, the greater water retention prop-
erties of the dung layers across abandoned pastoralist settlements
appears to play a critical role in enhancing survival rates during the
first and most critical dry season (Reid & Ellis 1995). Another, pos-
sibly beneficial consequence is that since the genus Acacia is very
effective for N2 symbiotic fixation in the soil (Cech et al. 2008: 992),
the presence of stands of A. tortilis on abandoned settlements may,
through leaching and ammonia volatilisation, help offset the loss of
N that otherwise occurs on such sites (Augustine 2003). As a result,
rings of A. tortilis, marking the location of former livestock pens,
are scattered across southern Turkana’s landscape. During the dry
season, these rings stand out as distinctive green patches in the land-
scape, and the history of human settlement is thus encoded within
the prevailing vegetation.
The age of the glades formed in this manner is variable, but
longitudinal studies, combined with oral information about the his-
tory of these sites, show that they are often older than 100 years
(e.g. Young et al. 1995; Augustine 2003; Muchiru et al. 2008).
Archaeologists familiar with these environments have also noted
that the glades often contain traces of previous settlement activity
of variable date, including both the Pastoral Iron Age (ca. 1200–800
BP) and the Pastoral Neolithic (ca. 4500–1200 BP), suggesting that
some glades have survived for considerable periods of time (Lane
2011). The lack of systematic investigation of the antiquity of spe-
cific glades, coupled with the well-documented evidence for frequent
localised shifting of vegetation patches within these landscapes
(Gillson 2004; Wiegand, Ward, & Saltz 2005) makes it difficult to
assess how long individual glades have existed. Recent research on
the Laikipia Plateau, in north-central Kenya, has, however, suggested

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
2
6
1

162 Eriksson et al.

a possibility that some glades can remain open for up to ca. 750 years
(Causey 2008; Lane 2011; Boles & Lane 2016) (Figures 6.1 and 6.2).
The presence of glades formed from abandoned pastoralist
settlements in the landscape also shapes the spatial patterning and
densities of grazers and browsers, which vary in response to the
changing vegetation mosaic. Thus, for instance, in Amboseli (south-
eastern Kenya), species such as zebra, wildebeest, Thomson’s gazelle,
Grant’s gazelle, onyx, and gerenuk all tend to concentrate around
abandoned settlements, while all other species, including livestock,
‘peak at intermediate distance from [abandoned] settlements and
then [decline] with distance away from settlements’ (Muchiru et al.
2008: 946). Moreover, aside from increasing overall density of wild
ungulates and the species richness of the surrounding vegetation at
the landscape scale, the localised concentration of nutrients in aban-
doned pastoralist settlements can have beneficial effects at other
trophic levels. For example, invertebrate populations, especially
flies and beetles, often show ‘significant positive correlations to [the]
amount of dung deposited on abandoned settlements’ (Söderström
& Reid 2010: 187). Larger populations of insects may also partly
account for the higher densities and wider species diversity of birds
near abandoned settlements (Morris, Western, & Maitumo 2009).
In summary, these diverse ecological relationships imply that
East Africa’s pastoralist landscapes, often tenuously envisaged as
wild spaces that have accommodated a human presence, are in fact
highly domesticated spaces in which the complex entanglements
among dung, humans, and a wide range of non-human ‘things’ have a
critical role to play (see Lane 2016, for further discussion).

6.5.2 Domesticated Landscapes in Europe


In large parts of Europe, semi-natural grassland ecosystems played
a central role in agricultural systems prior to the processes of mod-
ernisation (Emanuelsson 2009). As the food base for livestock hus-
bandry, as well as a nutrient base for cultivation, grasslands spatially
dominated most domesticated landscapes. Many studies have shown

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
3
6
1

Concepts for Integrated Research in Historical Ecology 163

Figure 6.4 Aerial photo from the province of Uppland, Sweden,


showing a mosaic of nature types formed by various combinations of
human impact and natural processes. In the front, grazing has shaped
a fodder-producing grassland out of a seashore reed belt, in which the
biomass-producing vegetation is influenced by both the grazing and
the water.
Photo: Bergslagsbild. Courtesy of the Uppsala County Administrative
Board and the Uppland Foundation.

that agricultural societies created grasslands by opening up wooded


landscapes, with the aim of obtaining productive grass swards, often
in combination with production of leaf fodder from scattered trees
and shrubs (see review in Eriksson & Cousins 2014). A wide range of
grassland biotopes were formed, the types depending on climatic and
geological prerequisites in combination with the land-use practices
of the local and regional agricultural societies (domesticated semi-
natural grasslands landscapes: Figures 6.4–6.6). Many aspects of the
ecology of grassland biotopes, and of the local grassland management,
affected each other reciprocally through feedback mechanisms.
Human-made grassland ecosystems are often referred to as
semi-natural biotopes. They are formed and maintained by anthropo-
genic disturbance regimes (mowing, grazing, burning, irrigating;

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
4
6
1

164 Eriksson et al.

Figure 6.5 Lybby village in the province of Närke, Sweden, showing a


pasture with pollarded and formerly pollarded trees.
Photo: Mårten Sjöbeck, 1934. Public domain.

Figure 6.6 Remnants of stone walls used for fencing out livestock from
crop fields and hay meadows (from the province of Halland, Sweden).
Photo: O. Eriksson.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
5
6
1

Concepts for Integrated Research in Historical Ecology 165

Gustavsson et al. 2011), but colonised predominantly by species per-


ceived as ‘wild’. In Scandinavia, these grassland systems probably
developed along with the introduction of agriculture around 6000
BP (Welinder 2011), but only became common much later, during
the early Iron Age around 2000 BP (Eriksson & Arnell 2017). Despite
repeated periods of decline and expansion (e.g. Lagerås 2007), these
grasslands dominated the landscape until a new agricultural system,
based on artificial fertilisers, was introduced between ca. 1860 and
1950 (Gadd 2011).
The use of grasslands by agrarian societies required adaptation,
primarily of practices of land use, which constitute the direct inter-
face between humans and the biotope. In addition, adaptation of the
entire local agricultural system was needed, in varying degrees. As a
result of the expansion of open grasslands, new agricultural regions
were created. During the more than two millennia that this grassland-
based system existed, local agricultural production systems changed
(Myrdal 2011; Pedersen & Widgren 2011), also affecting the distri-
bution and structure of grasslands and, as a consequence, their bio-
logical composition (Eriksson 2013; Eriksson & Cousins 2014). The
formation of domesticated landscapes based on a grassland ecosys-
tem thus led to an expansion of the realised niche of humans, parallel
to that of grassland species. The process of expansion of agricultural
communities by opening up the landscape and finding ways of using
different types of grassland and fields has been described in several
ethnological studies from Sweden (e.g. Atlestam 1942; Johansson
1947; Campbell 1948; and recently Kjellström 2012). These studies
show how the subsistence and structure of local communities devel-
oped over time, and how this process was affected by local ecological
conditions in combination with cultural and religious interaction,
markets, and the surrounding society.
Although none of the authors cited here uses the term niche,
their descriptions of the preconditions for these local, rural societies
are similar to the ecological definition of the realised niche of a spe-
cies, i.e. the set of environmental conditions actualised by a species

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
6
1

166 Eriksson et al.

at a specific time and place, allowing the species to persist. From


this perspective, ‘niche’ can be used for situating and understand-
ing human agency and behaviour in the environment, as elements
derived from ‘ecologies of practice’ (Nyerges 1997; Balée 2006). Such
a perspective is anthropogenic in focus, constituted by the role of
everyday human practice in shaping environments. Everyday human
ecological practices are structured by the social organisational prin-
ciples of land use, which in turn led to the shaping of domesticated
landscapes constituted by a variety of sometimes overlapping niches.
These can be considered to express landscape habitus, or the biocul-
tural heritage of socio-environmental interaction over time. Seen
from this perspective, landscapes are both the medium and the out-
come of human mind and agency in the environment. They also
provide a platform for formulating a conceptual framework that inte-
grates landscape ecology and social history. In a sense, humans have
constructed their own cultural niche, and simultaneously created
niches for various other species, both ‘wild’ and ‘domesticated’.

6.5.3 Domesticated Landscapes and Entanglement


The overall environment of local inhabitants consists of both
a biophysical and a societal environment – the latter including
socio-economic, cultural, and religious arenas (cf. Vestbö-Franzén
2005: 10). In a wider sense, this clearly makes the human niche a cul-
tural one (cf. Boyd et al. 2011). Practices and environmental states are
closely linked. On one hand, agrarian households created their own
domesticated environment, as seen in the examples from East Africa
and Europe. On the other hand, practices needed to be continuously
adjusted to shifting environmental conditions, thus linking the con-
cept of adaptation to local conditions.
It is here that the notion of entanglement, as explored in detail
by Hodder (2012: 179–205) with reference to the changing history of
human-thing relations at the large Neolithic and Chalcolithic settle-
ment of Çatalhöyük, Turkey (ca. 7500–5700 BC), offers a degree of
conceptual clarity. As he argues, individual and societal choices,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
7
6
1

Concepts for Integrated Research in Historical Ecology 167

even those aimed at meeting fairly mundane utilitarian needs, bind


together people and things (understood here to encompass elements
of the biophysical world as well as humanly manufactured mater-
ial culture) into relations that typically entrain subsequent choices
and actions. The intended and unintended consequences of differ-
ent interventions in and rearrangements of the order of humans and
things create new opportunities and challenges that, in turn, require
fresh intervention and rearrangement. Yet most human practice is
routine, undertaken in a manner that more or less replicates pre-
vious practice with the intention of reproducing the structure of
socio-ecological relations. This leads to a constant tension between,
on one hand, what humans want, and, on the other, what things
want (Gosden 2005). There is also a degree of mutual constitution,
in that the distinctiveness of each entity – humans, landscapes,
ecological niches, etc. – only emerges from the processes of their
entanglement. This is not through interaction, but instead – to use
the term of the theoretical physicist Karen Barad – via intra-action.
In Barad’s words (2007: 33), this means that ‘distinct entities do not
precede’ events or relations, ‘but rather emerge through, their intra-
actions’. Consequently, ‘agencies are only distinct in relation to their
mutual entanglement; they don’t exist as individual elements’ (Barad
2007: 33). One further implication of Barad’s notion of intra-action
is that relations, be they dialectical or otherwise, have no independ-
ent ontological existence, separate from that which they connect,
but emerge through the very processes of iterative intra-activity that
they enable. It is precisely this cumulative, performative, consti-
tuting, and unfolding process of ‘binding’ that Hodder’s concept of
entanglement is intended to convey. As such, although this concept
draws on ideas about assemblage, dialectical relations, and even the
affordances of ‘things’, in our view, it also goes beyond each of these
concepts and in many respects attempts to draw them together into
a more unifying analytical framework.
One particularly interesting way forward is to think of land-
scapes as entangled places, borrowing from Hodder’s (2012) concept,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
8
6
1

168 Eriksson et al.

in which humans, other species, and ‘things’ operate within pre-


existing relations of dependence and dependency. This is illustrated
by the example of the East African savanna landscapes inhabited by
pastoralist groups, that depend on livestock for their livelihood. The
livestock, in turn, depend on the pastoralists for their protection and
care, which involves them being kept overnight in pens at the centre
of the settlements. Regular penning of livestock inside settlement
enclosures creates dense accumulations of dung. After a number
of years, the dung and associated parasites make these settlements
uninhabitable, resulting in the abandonment and relocation of enclo-
sures. This recurrent settlement mobility is also necessitated by the
steady depletion of the grazing sward in the immediate vicinity of the
enclosure and the progressive removal of trees and bushes for use as
building materials and fuel. Thus, both humans and livestock are not
only dependent on another ‘thing’ – the settlement enclosure – but,
these relationships also establish other relationships of dependency.
The enclosure provides security – both from wild animals and
other, potentially more hostile groups of humans – but the actions of
humans and animals work against sustaining such secure places for
any length of time. At the same time, human dependency on other
things – grass, wood, forage, and so forth – reduces the inhabitabil-
ity of the space immediately surrounding the enclosure, resulting in
abandonment. Yet, as outlined earlier in this chapter, the act of aban-
donment initiates a new sequence of ecological relationships that, in
time, both literally and metaphorically revive the landscape by mak-
ing it re-inhabitable. In this way, the landscape is ‘restored’, initially
through a restricted number of non-human species, and subsequently
through a steadily widening – but still limited – range of species,
including humans and their livestock. Again, relations of dependence
and dependency among humans, non-human species, and things are
at play. Moreover, this recurrent process of abandonment and reoccu-
pation effectively works to reinsert pastoralists into their landscape
setting, thereby imbuing this with history, from which notions of
heritage value and a sense of place and belonging are derived (Straight

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
9
6
1

Concepts for Integrated Research in Historical Ecology 169

el al. 2016). In their inhabitation of these savanna landscapes, pas-


toralists also learn to ‘read’ this history of vegetation succession,
recognising the immediate, but nonetheless relative utilitarian
value of different ecological niches, while also remembering their
specific histories of human-thing, thing-thing, and human-human
entanglements.

6. 6 Concluding Remarks
This chapter demonstrates how a mutual borrowing of concepts can
be fruitful for research in both landscape ecology and social history.
Such mutual conceptual frames are crucial as they allow researchers
from different fields to communicate. They also promote the broad-
ening of scholarship by encouraging crossovers among the natural
sciences and humanities. A common conceptual base promotes dia-
logue on important issues such as assumptions, sources, and results,
thus creating a base for new research questions. It is clear that such
dialogues, mixing and synthesising perspectives, will ultimately gen-
erate new and unique insights.
Many of the issues that environmental management grapples
with today can only be answered through the long-term understand-
ing of both ecosystem dynamics (Willis et al. 2007; Seddon et al. 2014)
and landscape formation (e.g. Maley & Brenac 1998; Foster 2002;
Emanuelsson 2009; Vellend et al. 2013). One such example concerns
agricultural landscapes where a historical-ecological perspective is
essential for both conservation biology (Emanuelsson 2009; Eriksson
& Cousins 2014) and the development of sustainable rural economies
in a globalised world (e.g. Del Mármol & Vaccaro 2015). History is
an essential component for understanding how different types of
nature (biotopes) are formed, and on which land-use practices today’s
biotopes rely (e.g. Verheyen et al. 1999; Cousins & Eriksson 2002;
Dupouey et al. 2002; Hermy & Verheyen 2007; Eriksson & Cousins
2014). Historical data are also instrumental for revising landscape
histories, challenging the often expressed – not least among ecolo-
gists and conservation biologists – myth of a ‘pristine’ landscape and

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
0
7
1

170 Eriksson et al.

‘degradation’ narratives, by referring to the long transformation of


landscapes by humans and domesticated animals (e.g. Fairhead &
Leach 1996; Rackham 1996; McCann 1999; Vera 2000; Willis, Gillson,
& Brncic 2004; Josefsson et al. 2010; Harris 2012; Ellis et al. 2013).
In line with these trends, Higgs and colleagues (2014) argue
for a new role for history in restoration ecology. Instead of ‘classical’
ecological restoration, which they describe as attempting to return
ecosystem to their historic trajectories, they suggest that historical
ecology should be used as a guide rather than a template. This entails
that process has priority over structure and composition and, above
all, that new approaches are required for addressing cultural needs
and human livelihood. This emphasis requires conceptual tools for
the study of processes, incorporating aspects of both landscape ecol-
ogy and social history. In conclusion, it is clear that an integrative
framework based on concepts of niche construction, adaptation,
landscape domestication, and entanglement will meet this demand,
and therefore also promote basic and applied research in historical
ecology.

References
Anderson, K. (1997). A walk on the wild side: a critical geography of domestica-
tion. Progress in Human Geography, 21, 463–85.
Archibald, S. (2008). African grazing lawns – how fire, rainfall, and grazer numbers
interact to affect grass community status. Journal of Wildlife Management,
72, 492–501.
Augustine, D. J. (2003). Long-term, livestock mediated redistribution of nitrogen
and phosphorus in an East African savanna. Journal of Applied Ecology, 40,
137–49.
Augustine, D. J. & McNaughton, S. J. (2004). Regulation of shrub dynamics by
native browsing ungulates on East African rangeland. Journal of Applied
Ecology, 41, 45–58.
Atlestam, P-O. (1942). Bohusläns ljunghedar, en geografisk studie. Gothenburg:
Meddelanden från Göteborgs högskolas Geografiska institution 30.
Balée, W. (2006). The research program of historical ecology. Annual Review of
Anthropology, 35, 75–98.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
1
7

Concepts for Integrated Research in Historical Ecology 171

Barad, K. (2007). Meeting the Universe Halfway: Quantum Physics and the
Entanglement of Matter and Meaning. Durham, NC: Duke University Press.
Barrow, E. C. G. (1990). Usufruct rights to trees: the role of Ekwar in dryland cen-
tral Turkana, Kenya. Human Ecology, 18, 163–76.
Bennet, J. (2010). Vibrant Matter: A Political Ecology of Things. Durham,
NC: Duke University Press.
Berglund, B. E., ed. (1991). The cultural landscape during 6000 years in southern
Sweden. Ecological Bulletins, 41, 1–495.
Blench, R. (2007). The intertwined history of silk-cotton and baobab. In R. J.
Cappers, ed., Fields of Change: Progress in African Archaeobotany. Groningen
Archaeological Studies 5. Groningen: Barkhuis Publishing & Groningen
University Library, pp. 1–19.
Boles, J. C. & Lane, P. J. (2016). The green, green grass of home: an archaeo-ecological
approach to pastoralist settlement in central Kenya. Azania: Archaeological
Research in Africa, 51(4), 507–30. Doi: 10.1080/0067270X.2016.1249587.
Bourdieu, P. (1977). Outline of a Theory of Practice. Cambridge: Cambridge
University Press.
Boyd, R., Richerson, P. J., & Henrich, J. (2011). The cultural niche: why social
learning is essential for human adaptation. Proceedings of the National
Academy of Sciences USA, 108, 10918–25.
Broughton, J. M., Cannon, M. D., & Bartelink, E. J. (2010). Evolutionary ecology,
resource depression, and niche construction theory: applications to Central
California hunter-gatherers and Mimbres-Mogollon agriculturalists. Journal
of Archaeological Method and Theory, 17, 371–421.
Butzer, K. W. (1982). Archaeology as Human Ecology: Method and Theory for a
Contextual Approach. Cambridge: Cambridge University Press.
Butzer, K. W. (1996). Ecology in the long view: settlement histories, agrosystemic
strategies, and ecological performance. Journal of Field Archaeology, 23, 141–50.
Butzer, K. W. (2005). Environmental history in the Mediterranean world: cross-
disciplinary investigation of cause-and-effect for degradation and soil erosion.
Journal of Archaeological Science, 32, 1773–1800.
Campbell, Å. (1948). Från Vildmark till bygd. En Etnologisk undersökning av
nybyggarkulturen I Lappland före industrialismens genombrott. Landsmåls-
och Folkminnesarkivet, Uppsala, B:5.
Callon, M. & Law, J. (1995). Agency and the hybrid collectif. South Atlantic
Quaterly, 94, 481–507.
Causey, M. (2008). Delineating Pastoralist Behaviour and Long-Term
Environmental Change: A GIS Landscape Approach on the Laikipia Plateau,
Kenya, PhD Dissertation, School of Archaeology, University of Oxford.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
2
7
1

172 Eriksson et al.

Cech, P. G., Kuster, T., Edwards, P. J., & Venterink, H. O. (2008). Effects of her-
bivory, fire and N2-fixation on nutrient limitation in a humid African savan-
nah. Ecosystems, 11, 991–1004.
Chase, J. M. & Leibold, J. M. (2003). Ecological Niches: Linking Classical and
Contemporary Approaches. Chicago: University of Chicago Press.
Childe, V. G. (1928). The Most Ancient East: The Oriental Prelude to European
Prehistory. London: Kegan Paul.
Clement, R. C., Denevan, W. M., Heckenberger, M. J., Junqueira, A. B., Neves, E.
G., Teixeira, W. G., & Woods, W. I. (2015). The domestication of Amazonia
before European conquest. Proceeding of the Royal Society B, 282, 20150813.
Clements, F. (1916). Plant Succession. Washington, DC: Carnegie Institution.
Cousins, S. A. O. & Eriksson, O. (2002). The influence of management history
and habitat on plant species richness in a rural hemiboreal landscape, Sweden.
Landscape Ecology, 17, 517–29.
Cowles, H. C. (1899). The ecological relations of the vegetation on the sand dunes
of Lake Michigan. Botanical Gazette, 27, 95–117, 167–202, 281–308, 361–91.
Crumley, C. (1987). Celtic settlement before the Conquest: the dialectics of
landscape and power. In C. L. Crumley & W H. Marquardt, eds., Regional
Dynamics: Burgundian Landscapes in Historical Perspective. San Diego,
CA: Academic Press, pp. 403–29.
Crumley, C. L. (ed.). (1994). Historical Ecology: Cultural Knowledge and Changing
Landscapes. Santa Fe, NM: School of American Research Press.
Crumley, C. L. (2007). Historical ecology: integrated thinking at multiple tempo-
ral and spatial scales. In A. Hornborg & C. Crumley, eds., The World System
and the Earth System: Global Socioenvironmental Change and Sustainability
since the Neolithic. Walnut Creek, CA: Left Coast Press, pp. 15–28.
Danchin, E. (2013). Avatars of information: towards an inclusive evolutionary syn-
thesis. Trends in Ecology and Evolution, 28, 351–8.
Darwin, C. R. (1859). On the Origin of Species by means of Natural Selection,
or the Preservation of Favoured Races in the Struggle for Life. London: John
Murray.
Davies, M. I. J. (2013). Environment in North American and European archaeology.
In M. I. J. Davies & F. N. M’Mbogori, eds., Humans and the Environment: New
Archaeological Perspectives for the 21st Century. Oxford: Oxford University
Press, pp. 3–25.
Deleuze, G. & Guattari, F. [1980] (2010). A Thousand Plateaus. London: Bloomsbury
Academic.
Del Mármol, C. & Vaccaro, I. (2015). Changing ruralities: between abandonment
and redefinition in the Catalan Pyrenees. Anthropological Forum, 25, 21–41.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
3
7
1

Concepts for Integrated Research in Historical Ecology 173

Denevan, W. M. (2011). The ‘pristine myth’ revisited. Geographical Review, 101,


576–91.
Dieter, J. (2012). Evolution and the paleolithic. Notae Praehistoricae, 32, 257–87.
Dobney, K. & Larson, G. (2006). Genetics and animal domestication: new win-
dows on an elusive process. Journal of Zoology, 269, 261–71.
Dunnell, R. C. (1971). Systematics in Prehistory. New York: Free Press.
Dupouey, J.-L., Dambrine, E., Lafitte, J. D., & Moares, C. (2002). Irreversible
impact of past land use on forest soils and biodiversity. Ecology, 83, 2978–84.
Duvall, C. S. (2007). Human settlement and baobab distribution in south-western
Mali. Journal of Biogeography, 34, 1947–61.
Earle, T. & Spriggs, M. (2015). Political economy in prehistory: a Marxist approach
to Pacific sequences. Current Anthropology, 56, 515–44.
Egerton, F. N. (1973). Changing concepts of the balance of nature. Quarterly
Review of Biology, 48, 322–50.
Ellen, R. (1982). Environment, Subsistence and System: The Ecology of Small-
Scale Social Formations. Cambridge: Cambridge University Press.
Ellis, E. C., Kaplan, J. O., Fuller, D. Q., Vavrus, S., Goldewijk, K. K., & Verburg, P.
H. (2013). Used planet: a global history. Proceedings of the National Academy
of Sciences USA, 110, 7978–85.
Ellis, E. C. & Ramankutty, N. (2008). Putting people in the map: anthropogenic
biomes of the world. Frontiers in Ecology and the Environment, 6, 439–47.
Emanuelsson, U. (2009). The Rural Landscape of Europe: How Man Has Shaped
European Nature. Stockholm: The Swedish Research Council Formas.
Erickson, C. L. (2006). The domesticated landscapes of the Bolivian Amazon. In
W. Balée & C. L. Erickson, eds., Time and Complexity in Historical Ecology.
New York: Columbia University Press, pp. 235–78.
Erickson, C. L. (2008). Amazonia: the historical ecology of a domesticated land-
scape. In H. Silverman & W. H. Isbell, eds., Handbook of South American
Archaeology. New York: Springer, pp. 157–83.
Erickson, C. L. (2010). The transformation of environment into landscape: the
historical ecology of monumental earthwork construction in the Bolivian
Amazon. Diversity, 2, 618–52.
Eriksson, O. (2013). Species pools in cultural landscapes – niche construction, eco-
logical opportunity and niche shifts. Ecography, 36, 403–13.
Eriksson, O. & Arnell, M. (2017). Niche construction, entanglement and land-
scape domestication in Scandinavian infield systems. Landscape Research,
42(1), 78–88. Doi: 10.1080/01426397.2016.1255316.
Eriksson, O. & Cousins, S. A. O. (2014). Historical landscape perspectives on
grasslands in Sweden and the Baltic region. Land, 3, 300–21.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
4
7
1

174 Eriksson et al.

Escobar, A. (1996). Construction nature: elements for a post structuralist political


ecology. Futures, 28, 325–43.
Fairhead, J. & Leach, M. (1996). Misreading the African Landscape: Society and
Ecology in a Forest-Savanna Mosaic. Cambridge: Cambridge University
Press.
Forman, R. T. T. & Godron, M. (1986). Landscape Ecology. New York: Wiley.
Foster, D. R. (2002). Thoreau’s country: a historical-ecological perspective on con-
servation in a New England landscape. Journal of Biogeography, 29, 1537–55.
Foucault, M. (1970). The Order of Things. New York: Random House.
Fraser, J. A., Frausin, V., & Jarvis, A. (2015). An intergenerational transmission of
sustainability? Ancestral habitus and food production in a traditional agro-
ecosystem of the Upper Guinea Forest, West Africa. Global Environmental
Change, 31, 226–38.
Frausin, V., Fraser, J. A., Narmah, W., Lahai, M. K., Winnebah, T. R., Fairhead, J.,
& Leach, M. (2014). ‘God made the soil, but we made it fertile’: gender, knowl-
edge, and practice in the formation and use of African Dark Earths in Liberia
and Sierra Leone. Human Ecology, 42, 695–710.
Fried, M. (1967). The Evolution of Political Society: An Essay in Political Economy.
New York: Random House.
Friedman, J. & Rowlands, M. (1978). Notes toward an epigenetic model of the
evolution of ‘civilization’. In J. Friedman & M. Rowlands, eds., The Evolution
of Social Systems. London: Duckworth, pp. 201–76.
Gadd, C. J. (2011). The agricultural revolution in Sweden 1700–1870. In J. Myrdal
& M. Morell, eds., The Agrarian History of Sweden from 4000 BC to AD 2000.
Lund: Nordic Academic Press, pp. 118–64.
Gero, J. & Conkey, W., eds. (1991). Engendering Archaeology. Oxford: Blackwell.
Giddens, A. (1984). The Constitution of Society: Outline of the Theory of
Structuration. Cambridge: Polity Press.
Gillson, L. (2004). Testing non-equilibrium theories in savannas: 1400 years of
vegetation change in Tsavo National Park, Kenya. Ecological Complexity, 1,
281–98.
Gilman, A. (1981). The development of social stratification in Bronze Age Europe.
Current Anthropology, 22, 1–24.
Gleason, H. A. (1926). The individualistic concept of the plant association.
Bulletin of the Torrey Botanical Club, 53, 7–26.
Godsoe, W. (2010). I can’t define the niche but I know it when I see it: a formal
link between statistical theory and the ecological niche. Oikos, 119, 53–60.
Gosden, C. (2005). What do objects want? Journal of Archaeological Method and
Theory, 12, 193–211.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
5
7
1

Concepts for Integrated Research in Historical Ecology 175

Gustavsson, E., Dahlström, A., Emanuelsson, M., Wissman, J., & Lennartsson,
T. (2011). Combining historical and ecological knowledge to optimise bio-
diversity conservation in semi-natural grasslands. In J. L. Pujol, ed., The
Importance of Biological Interactions in the Study of Biodiversity. New York:
InTech Publishers, pp. 173–96.
Haraway, D. (1991). Simians, Cyborgs, and Women: The Reinvention of Nature.
London: Free Association Books.
Harris, D. (1996). Domesticatory relationships of people, plants and animals.
In R. Ellen & K. Fukui, eds., Redefining Nature: Ecology, Culture and
Domestication. Oxford: Berg, pp. 437–63.
Harris, S. E. (2012). Cyprus as a degraded landscape or resilient environment in the
wake of colonial intrusion. Proceedings of the National Academy of Sciences
USA, 109, 3670–5.
Hermy, M. & Verheyen, K. (2007). Legacies of the past in the present-day forest
biodiversity: a review of past land-use effects on forest plant species composi-
tion and diversity. Ecological Research, 22, 361–71.
Higgs, E., Falk, D. A., Guerrini, A., Hall, M., Harris, J., Hobbs, R. J., Jackson, S. T.,
Rhemtulla, J. M., & Throop, W. (2014). The changing role of history in restora-
tion ecology. Frontiers in Ecology and the Environment, 12, 499–506.
Hinchcliffe, S. (2007). Geographies of Nature: Societies, Environments, Ecologies.
Los Angeles: Sage Publications.
Hodder, I. (1982). Theoretical archaeology: a reactionary view. In I. Hodder, ed.,
Symbolic and Structural Archaeology. Cambridge: Cambridge University
Press, pp. 1–16.
Hodder, I. (2012). Entangled: An Archaeology of the Relationships between
Humans and Things. Oxford: Blackwell Wiley.
Hubbell, S. P. & Foster, R. S. (1986). Biology, chance, and history and the struc-
ture of tropical rainforest tree communities. In J. Diamond & J. D. Case, eds.,
Community Ecology. New York: Harper and Row, pp. 314–43.
Hutchinson, G. E. (1957). Concluding remarks. Cold Spring Harbor Symposium of
Quantitative Biology, 22, 415–27.
Jarman, M. R. & Wilkinson, P. F. (1972). Criteria of animal domestication. In E. S.
Higgs, ed., Papers in Economic Prehistory. Cambridge: Cambridge University
Press, pp. 83–96.
Johansson, L. (1947). Bebyggelse och folkliv i det gamla Frostviken. Landsmåls-
och Folkminnesarkivet i Uppsala B:3. Uppsala.
Josefsson, T., Gunnarsson, B., Liedgren, L., Bergman, I., & Östlund, L. (2010).
Historical human influence on forest composition and structure in boreal
Fennoscandia. Canadian Journal of Forest Research, 40, 872–84.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
6
7
1

176 Eriksson et al.

Kareiva, P., Watts, S., McDonald, R., & Boucher, T. (2007). Domesticated
nature: shaping landscapes and ecosystems for human welfare. Science, 316,
1866–9.
Kendal, J., Tehrani, J. J., & Odling-Smee, J. (2011). Human niche construction in
interdisciplinary focus. Philosophical Transactions of the Royal Society B,
366, 785–92.
Kirch, P. V. (1984). The Evolution of Polynesian Chiefdoms. Cambridge: Cambridge
University Press.
Kjellström, R. (2012). Nybyggarliv i Vilhelmina. 1. Träd och växter som resurs.
Acta Academiae Regiae Gustavi Adolphi 119, CBM Publications 65. Uppsala.
Kricher, J. (2009). The Balance of Nature: Ecology’s Enduring Myth. Princeton,
NJ: Princeton University Press.
Lagerås, P. (2007). The Ecology of Expansion and Abandonment: Medieval
and Post-medieval Land-Use and Settlement Dynamics in a Landscape
Perspective. Stockholm: National Heritage Board, Sweden.
Laland, K. N. & O’Brien, M. J. (2011). Cultural niche construction: an introduc-
tion. Biological Theory, 6, 191–202.
Lamprey, H. F., Halevy, G., & Makacha, S. (1974). Interactions between Acacia
bruchid seed beetles and large herbivores. East African Wildlife Journal,
12, 81–5.
Lane, P. J. (2011). An outline of the later Holocene archaeology and precolonial
history of the Ewaso Basin, Kenya. Smithsonian Contributions to Zoology,
632, 11–30.
Lane, P. J. (2016). Entangled banks and the domestication of East African pastoral-
ist landscapes. In F. Fernandini & L. Der, eds., Archaeology of Entanglement.
Walnut Creek, CA: Left Coast Press, pp. 127–50.
Latour, B. (1999). Pandora’s Hope: Essays in the Reality of Science Studies.
Cambridge, MA: Harvard University Press.
Latour, B. (2000). When things strike back – a possible contribution of science
studies to the social sciences. British Journal of Sociology, 51, 107–23.
Latour, B. (2005). Reassembling the Social: An Introduction to Actor-Network-
Theory. Oxford: Oxford University Press.
Law, J. & Mol, A. (2002). Complexities: Social Studies of Knowledge Practices.
Durham, NC: Duke University Press.
Leach, H. (2003). Human domestication reconsidered. Current Anthropology, 44,
349–68.
Maley, J. & Brenac, P. (1998). Vegetation dynamics, palaeoenvironments and cli-
matic changes in the forests of western Cameroon during the last 28,000 years
B.P. Review of Palaeobotany and Palynology, 99, 157–87.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
7
1

Concepts for Integrated Research in Historical Ecology 177

Maranz, S. (2009). Tree mortality in the African Sahel indicates an anthropogenic


ecosystem displaced by climate change. Journal of Biogeography, 36, 1181–93.
Marcus, G. A. & Fischer, M. F., eds. (1986). Anthropology as Cultural Critique: An
Experimental Moment in the Human Sciences. Chicago: University of
Chicago Press.
Marquardt, W. H. (1992). Dialectical archaeology. In M. B. Schiffer, ed.,
Archaeological Method and Theory, Volume 4. Tucson: University of Arizona
Press, pp. 101–40.
Mather, C. (2003). Shrines and the domestication of landscape. Journal of
Anthropological Research, 59, 23–45.
McCann, J. (1999). Green Land, Brown Land, Black Land: An Environmental
History of Africa, 1800–1990. London: James Currey.
Mesoudi, A. (2011). Cultural Evolution: How Darwinian Theory Can Explain
Human Culture and Synthesize the Social Sciences. Chicago: University of
Chicago Press.
Meyer, W. J. & Crumley C. L. (2011). Historical ecology: using what works to cross
the divide. Atlantic Europe in the first millennium BC: crossing the divide.
In T. Moore & L. Armada, eds., Atlantic Europe in the First Millennium
BC: Crossing the Divide. Oxford: Oxford University Press, pp. 109–34.
Morris, D. L., Western, D., & Maitumo, D. (2009). Pastoralist’s livestock and set-
tlements influence game bird diversity and abundance in a savanna ecosystem
of southern Kenya. African Journal of Ecology, 47, 48–55.
Muchiru, A. N., Western, D. J., & Reid, R. S. (2008). The role of abandoned pas-
toral settlements in the dynamics of African large herbivore communities.
Journal of Arid Environments, 72, 940–52.
Myrdal, J. (2011). Farming and feudalism. In J. Myrdal & M. Morell, eds., The
Agrarian History of Sweden from 4000 BC to AD 2000. Lund: Nordic
Academic Press, pp. 72–117.
Nyerges, A. E. ed. (1997). The Ecology of Practice: Studies of Food Crop Production
in Sub-Saharan West Africa. Amsterdam: Gordon & Breach.
O’Brien, M. J. & Laland, K. N. (2012). Genes, culture, and agriculture: an example
of human niche construction. Current Anthropology, 53, 434–70.
O’Connor, T. P. (1997). Working at relationships: another look at animal domesti-
cation. Antiquity, 71, 149–56.
Odling-Smee, J., Erwin, D., Palkovacs, E. P., Feldman, M. W., & Laland, K. N.
(2013). Niche construction theory: a practical guide for ecologists. Quarterly
Review of Biology, 88, 3–28.
Odling-Smee, F. J., Laland, K. N., & Feldman, M. W. (2003). Niche Construction: The
Neglected Process in Evolution. Princeton, NJ: Princeton University Press.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
8
7
1

178 Eriksson et al.

Odum H. T. (1983). Systems Ecology: An Introduction. New York: Wiley.


O’Neill, R. V., DeAngelis, D. L., Waide, J. B., & Allen, T. F. H. (1986). A Hierarchical
Concept of Ecosystems. Princeton, NJ: Princeton University Press.
Pearman, P. B., Guisan, A., Broennimann, O., & Randin, C. F. (2008). Niche
dynamics in space and time. Trends in Ecology and Evolution, 23, 149–58.
Pedersen E. A. & Widgren, M. (2011). Agriculture in Sweden, 800 BC–AD 1000. In
J. Myrdal & M. Morell, eds., The Agrarian History of Sweden from 4000 BC to
AD 2000. Lund: Nordic Academic Press, pp. 46–71.
Pickett, S. T. A. & White, P. S. (1985). Patch dynamics: a synthesis. In S. T. A.
Pickett & P. S. White, eds., The Ecology of Natural Disturbance and Patch
Dynamics. San Diego, CA: Academic Press, pp. 371–84.
Pinker, S. (2010). The cognitive niche: coevolution of intelligence, sociality,
and language. Proceedings of the National Academy of Sciences USA, 107,
8993–9.
Post, D. M. & Palkovacs, E. P. (2009). Eco-evolutionary feedbacks in community and
ecosystem ecology: interactions between the ecological theatre and the evolu-
tionary play. Philosophical Transactions of the Royal Society B, 364, 1629–40.
Rackham, O. (1996). Ecology and pseudo-ecology: the example of ancient Greece.
In G. Shipley, ed., Human Landscapes in Classical Antiquity: Environment
and Culture. London and New York: Routledge, pp. 16–43.
Rappaport, R. A. (1968). Pigs for the Ancestors: Ritual in the Ecology of a New
Guinea People. New Haven, CT: Yale University Press.
Reid, R. S. & Ellis, J. E. (1995). Impacts of pastoralists on woodlands in South
Turkana, Kenya: livestock-mediated tree recruitment. Ecological Applications,
5, 978–92.
Renfrew, C. (1972). The Emergence of Civilization: The Cyclades and the Aegean
in the Third Millennium B.C. London: Methuen.
Renfrew, C., Rowlands, M. J., & Segraves, B. A. eds. (1982). Theory and Explanation
in Archaeology: The Southampton Conference. London: Academic Press.
Rindos, D. (1984). The Origins of Agriculture: An Evolutionary Perspective.
London: Academic Press.
Robbins, P. (2012). Political Ecology, 2nd ed. Oxford: Wiley-Blackwell.
Rosman, A. & Rubel, P. G. (1989). Stalking the wild pig: hunting and horticulture in
Papua New Guinea. In S. Kent, ed., Farmers as Hunters. Cambridge: Cambridge
University Press, pp. 27–36.
Russell, N. (2002). The wild side of animal domestication. Society & Animals,
10, 285–302.
Sahlins, M. (1985). Islands of History. Chicago: University of Chicago Press.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
9
7
1

Concepts for Integrated Research in Historical Ecology 179

Seddon, A. W. R., Machay, A. W, Baker, A. G., & Birks, H. G. (2014). Looking


forward through the past: identification of 50 priority research questions in
palaeoecology. Journal of Ecology, 102, 256–67.
Shanks, M. & Tilley, C. (1987). Re-constructing Archaeology: Theory and Practice.
Cambridge: Cambridge University Press.
Smith, B. D. (2007). Niche construction and the behavioral context of plant and
animal domestication. Evolutionary Anthropology, 16, 188–99.
Smith, B. D. (2011a). A cultural niche construction theory of initial domestica-
tion. Biological Theory, 6, 260–71.
Smith, B. D. (2011b). General patterns of niche construction and the management
of ‘wild’ plant and animal resources by small-scale pre-industrial societies.
Philosophical Transactions of the Royal Society B, 366, 836–48.
Soberón, J. (2007). Grinnellian and Eltonian niches and geographic distributions of
species. Ecology Letters, 10, 1115–23.
Söderström, B. & Reid, R. S. (2010). Abandoned pastoral settlements provide con-
centrations of resources for savannah birds. Acta Oecologica, 36, 184–96.
Spector, J. D. (1993). What this Awl Means: Feminist Archaeology at a Wahpeton
Dakota Village. Minneapolis: Minnesota Historical Society Press.
Spriggs, M., ed. (1984). Marxist Perspectives in Archaeology. Cambridge: Cambridge
University Press.
Straight, B., Lane, P. J., Hilton, C., & Letua, M. (2016). ‘Dust People’: Samburu
perspectives on disaster, identity, and landscape. Journal of Eastern African
Studies, 10, 168–88.
Szabó, P. (2010). Why history matters in ecology: an interdisciplinary perspective.
Environmental Conservation, 37, 380–7.
Szabó, P., & Hedl, R. (2011). Advancing the integration of history and ecology for
conservation. Conservation Biology, 25, 680–7.
Terrell, J. E. & Hart, J. O. (2006). Domesticated landscapes. In B. David & J.
Thomas, eds., Handbook of Landscape Archaeology. Walnut Creek, CA: Left
Coast Press, pp. 328–32.
Thompson, J. N. (1994). The Coevolutionary Process. Chicago: University of
Chicago Press.
Thrift, N. (2008). Non-representational Theory: Space, Politics, Affect. London
and New York: Routledge.
Tooby, J. & DeVore, I. (1987). The reconstruction of hominid behavioral evolution
through strategic modeling. In W. G. Kinzey, ed., The Evolution of Human
Behavior: Primate Models. Albany: State University of New York Press,
pp. 183–237.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
0
8
1

180 Eriksson et al.

Vayda, A. P., ed. (1969). Environment and Cultural Behavior. New York: The
Natural History Press.
Vellend, M., Brown, C. D., Kharouba, H. M., McCune, J. L., & Myers-Smith, I.
H. (2013). Historical ecology: using unconventional data sources to test for
effects of global environmental change. American Journal of Botany, 100,
1294–1305.
Vera, F. W. M. (2000). Grazing Ecology and Forest History. Wallingford: CABI
Publishing.
Verheyen, K., Bossuyt, B., Hermy, M., & Tack, G. (1999). The land use history
(1278–1990) of a mixed hardwood forest in western Belgium and its relation-
ship with chemical soil characteristics. Journal of Biogeography, 26, 1115–28.
Vestbö-Franzén, A. (2005). Råg och rön: om mat, människor och landskaps-
förändringar i norra Småland, ca 1550–1700. Dept. of Cultural Geography,
Stockholm University 132.
Warming, E. (1895). Plantesamfund. Kjøbenhavn: P.G. Philipsens Forlag. (English
edition: Oecology of Plants. 1909. Oxford: Clarendon Press).
Watson, P. J., LeBlanc, S. A., & Redman, C. L. (1971). Explanation in Archeology: An
Explicitly Scientific Approach. New York: Columbia University Press.
Watt, A. S. (1947). Pattern and process in the plant community. Journal of Ecology,
35, 1–22.
Welinder, S. (2011). Early farming households. In J. Myrdal & M. Morell, eds.,
The Agrarian History of Sweden from 4000 BC to AD 2000. Lund: Nordic
Academic Press, pp. 18–45.
Whatmore, S. (2002). Hybrid Geographies: Natures, Culture, Spaces. London: Sage
Publications.
Whiten, A. & Erdal, D. (2012). The human socio-cognitive niche and its evolution-
ary origins. Philosophical Transactions of the Royal Society B, 367, 2119–29.
Whittaker, R. H. (1975). Communities and Ecosystems. New York: Macmillan.
Widgren, M. (2012). Landscape research in a world of domesticated landscapes: the
role of values, theory, and concepts. Quaternary International, 251, 117–24.
Wiegand, K., Ward, D., & Saltz, D. (2005). Multi-scale patterns and bush encroach-
ment in an arid savanna with a shallow soil layer. Journal of Vegetation
Science, 16, 311–20.
Willis, K. J., Araújo, M. B., Bennett, K. D., Figueroa-Rangel, B., Froyd, C. A., &
Myers, N. (2007). How can a knowledge of the past help to conserve the
future? Biodiversity conservation and the relevance of long-term ecological
studies. Philosophical Transactions of the Royal Society B, 362, 175–86.
Willis, K. J., Gillson, L., & Brncic, T. M. (2004). How ‘virgin’ is virgin rainforest?
Science, 304, 402–3.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
1
8

Concepts for Integrated Research in Historical Ecology 181

Wilson, D. S. (2005). Evolutionary social constructivism. In J. Gottschall & D. S.


Wilson, eds., The Literary Animal: Evolution and the Nature of Narrative.
Evanston, IL: Northwestern University Press, pp. 20–37.
Wolch, J. R. & Emel, J., eds. (1998). Animal Geographies: Place, Politics, and
Identity in the Nature-Culture Borderlands. London: Verso.
Wolf, E. R. (1982). Europe and the People without History. Berkeley: University
of California Press.
Worster, D. (1994). Natures Economy, a History of Ecological Ideas, 2nd ed.
Cambridge: Cambridge University Press.
Wu, J. & Loucks, O. (1995). From balance of nature to hierarchical patch dynam-
ics: a paradigm shift in ecology. Quarterly Review of Biology, 70, 439–66.
Wylie, A. (1985). Putting Shakertown back together: critical theory in archaeol-
ogy. Journal of Anthropological Archaeology, 4, 133–47.
Young, T. P., Partridge, N., & Macrae, A. (1995). Long-term glades in Acacia bush-
land and their edge effects in Laikipia, Kenya. Ecological Applications, 5,
97–108.
Zeder, M. A. (2011). The origins of agriculture in the Near East. Current
Anthropology, 52, S221–S235.
Zeuner, F. E. (1963). A History of Domesticated Animals. London: Hutchinson.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:43, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.006
2
8
1

7 Diversity in Ecological
and Social Contexts
Tommy Lennartsson, Ove Eriksson, Anamaria
Iuga, Jesper Larsson, Jon Moen, Michael D.
Scholl, Anna Westin, and Carole L. Crumley

7.1 Introduction
The holistic perspective of historical ecology places particular
emphasis on bridging the persisting rift between the Earth system
sciences, the social sciences, and humanities (Snow 1959). Much
has been written about how relatively recent events and ideas of the
twentieth century have estranged scholarly communities from one
another, resulting in an incomplete understanding of complex inter-
action among people, other life forms, and things (Hodder 2012; see
also Chapter 6).
This is not to imply that molecular biologists, Icelandic-saga
specialists, or cognitive psychologists do not know a lot about their
own subjects; close, empirical scrutiny is the foundation of inductive
science. In contrast, inductive reasoning creates connections to other
fields that are logically relevant to the study of both environmental
and societal change, and it can be questioned how we can study land-
use history without a wide variety of supporting disciplines (Bloch
2015). Each discipline employs methods and concepts that have
advanced its specific research needs. By reframing for different needs
and adopting a more accommodating approach to the requirements
of other research areas, the different disciplines can help to create
more comprehensive knowledge. The aim of this chapter is therefore
to examine how scholars can collaborate, while also continuing with
their own specialised work.
Chapter 8 investigates the interpersonal and group dynamics
of collaboration, as well as the key ethical, tactical, and structural
aspects. This chapter explores how theoretical issues arising during
182

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
3
8
1

Diversity in Ecological and Social Contexts 183

the course of such collaborations may be pursued. We begin by exam-


ining the use of the term diversity in both biological disciplines and
in social sciences/humanities, including a discussion of similar-
ities and differences between conventional biological and cultural
perspectives of diversity, and how these perspectives can benefit
from each other. The latter part of this chapter examines the hybrid
term biocultural diversity, and this term is used to discuss human-
environment interaction, partly through four case studies.

7.1.1 Terminology
For the purpose of this chapter, the term cultural diversity denotes
all aspects of diversity related to societies, e.g. human practices,
knowledge, traditions, language, and organisation, while biological
diversity denotes the diversity of ecosystems and organisms. Also
our use of the terms humanity, nature, and environment refers to
these two aspects of the Biosphere. We acknowledge the millennia-
long ontological, epistemological, and methodological discussion of
perceptions of ‘culture’ and ‘nature’ (e.g. Glacken 1967; Descola &
Pálsson 1996), including the potentially differing perspectives of the
scholars who observe and define nature-culture relationships and the
local communities that are part of these relationships (Couture 2000;
Johnson & Murton 2007). The aim of this chapter is, however, not to
demarcate the boundary between culture and nature, but to explore
the zone where the two merge. Our use of the term biological diver-
sity is the definition adopted by the Convention on Biodiversity,1
and we employ the UNESCO universal declaration on cultural diver-
sity: ‘diversity of cultures’ is the ‘co-existence of a difference in behav-
iour, traditions and customs’.2 Maffi and Woodley (2010: 4) describe
cultural diversity as ‘the diversity of the world’s cultural systems’,
which, just like our use of the term, accommodates the diversity of
cultural expressions, both between and within societies.

1
www.cbd.int/convention/text/, Article 2.
2
www.unesco.org/new/en/social-and-human-sciences/themes/international-
migration/glossary/cultural-diversity/

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
4
8
1

184 Lennartsson et al.

7.2 Biological Diversity


It was in ecology that the concept of diversity was first explicitly
developed and became a specialised field of research. Ecological
research examines both causes and effects of diversity in ecologi-
cal systems, as well as the important field of biodiversity loss. The
simplest measure of biological diversity is species richness, and in
order to compensate for differences in the abundance of different spe-
cies, several heterogeneity indices have been developed. In order to
link diversity to ecosystem, the concepts of food web diversity (Paine
1966), diversity of ecological functional groups (e.g. Root 1967), and
diversity of key species (Paine 1969) have been used.

7.2.1 Causes of Biological Diversity


In early research, biologists noticed geographical gradients of diver-
sity, between, e.g. the tropics and temperate regions (Wallace 1878),
different altitudes (von Humboldt & Bonpland 1807), islands of dif-
ferent sizes, and degrees of isolation (MacArthur & Wilson 1967),
as well as mountain ranges and peninsulas (Simpson 1964). Several
explanations have been suggested for the observed patterns, of which
some may be particularly relevant for biocultural diversity:

• Ecological age, allowing for species’ colonisation of habitats, and


evolutionary age, allowing for genetic diversification of plants and
animals (e.g. Sanders 1968).
• Predictability, stability, and continuity of the environment are assumed
to increase diversity, as seen, e.g. in the differences between the tropics
and temperate regions subject to periods of glaciation (e.g. Fischer 1960).
Periodic changes in the Earth’s climate are seen to decrease gradual
speciation and specialisation of species, but also to increase species’
distribution ranges and speciation by abrupt mechanisms, and to favour
dispersibility (Dynesius & Jansson 2000).
• The combined effect of food availability, water availability, and
temperature as a mechanism for, e.g. the tropics-temperate gradient
(Hawkins et al. 2003).
• Seasonality may reduce species richness either by killing species
dependent on benign environmental conditions (von Humboldt 1808), or

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
5
8
1

Diversity in Ecological and Social Contexts 185

by reducing the overall annual availability of food, water, and sufficient


temperature (Hawkins et al. 2003).
• Competition for resources may increase diversity by enhancing the
subdivision of resources into smaller units and also evolution of
specialised strategies (Menge & Sutherland 1976).
• Spatial heterogeneity influences how diversity on a landscape scale
(gamma diversity) is composed of diversity between (beta diversity)
and within (alfa diversity) habitats or patches (Whittaker 1960, 1972;
MacArthur 1965).
• Predation, herbivory, or other types of disturbance and stress may
increase diversity by reducing the dominance of a few competitive species
(Paine 1966; Harper 1969).

7.2.2 Effects of Biological Diversity


Ecological research pays considerable attention to the scalar effects
of biological diversity, especially whether diversity increases the sta-
bility and/or resilience of landscapes, ecosystems, and populations.
Stability is the result of resistance, i.e. the capacity to minimise
fluctuations (Lewontin 1969), whereas resilience is the capacity of
a defined system to persist in spite of fluctuations (Holling 1973).
When using either of these terms, it is necessary to define their con-
text, i.e. stability/resilience of ‘what’ under impact of ‘what’?
The idea that biological diversity results in stability was intro-
duced by researchers in the 1950s (e.g. MacArthur 1955; Elton 1958;
see McNaughton 1977). Subsequent research has found correlations
between species richness and certain aspects of stability in ecological
systems, in particular when using a single-stability measure, usually
biomass production, which represents an aggregate of the entire sys-
tem (e.g. Hector et al. 2010). The mechanisms determining correla-
tions between species richness and stability are still unclear. Buffering
through asynchronous responses of different species to environmen-
tal dynamics seems to be a general mechanism (Gonzales & Loreau
2009). It is not known, however, if species richness enhances such
buffering due to the presence of many complementary species, or if it
is related to the increased chance of recruitment of a few key species

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
6
8
1

186 Lennartsson et al.

(Loreau et al. 2001). Some of the observed effects may be the result of
the statistical methods used (Doak et al. 1998).

7.3 Cultural Diversity


The British anthropologist E. B. Tylor (1871: 1) was the first to use
the term culture as a general descriptor of the diverse social frame-
works that characterise humanity: ‘Culture . . . is that complex whole
which includes knowledge, belief, art, morals, law, custom, and any
other capabilities and habits acquired by man as a member of soci-
ety.’ While there are many definitions of culture, it is generally seen
as the shared knowledge or ‘world view’ of a specific group of humans
and passed on across generations. Culture is symbolic: its meanings
and applications are learned and transmitted, upheld and revised, and
are always framed by their material and cognitive contexts.
Cultural diversity evolves through individual and collective
innovations that are shaped by the circumstances in which they are
produced, as well as rules and norms. Cultural change is dialectical:
it is driven by changing surroundings that reshape traditions, val-
ues, memories, and experiences that in turn modify human thought
and activity and ultimately the environment. The broad diversity of
human society thus constitutes a rich source for new ideas in a rap-
idly changing world. Linguistic diversity forms a rich source mate-
rial for accumulated knowledge about places its speakers inhabit.
The many religions and belief systems of the world transmit values,
ethics, and place-based stewardship across generations. Diversity in
financial, political, social, and other terms of human welfare serves
as a reminder that high diversity is not necessarily a desired state.
Not all types of cultural diversity are related to biological diver-
sity: such relationships are mainly found when cultural diversity is
linked to the use of natural resources. Globally, diversity of the basic
ways of making a living is often successfully practised in marginal
environments, for example by combining animal husbandry, fishing,
and agriculture (Panjek & Larsson 2017). This has been termed diver-
sification of livelihood by Ellis (1998), who defines it as: ‘the process

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
7
8
1

Diversity in Ecological and Social Contexts 187

by which rural families construct a diverse portfolio of activities and


social support capabilities’. The choices leading to the adoption of a
particular livelihood include economic, practical, and ecological con-
cerns directly linked to the activities required for earning the liveli-
hood, as well as cultural concerns, which may be strongly but more
indirectly related to the livelihood strategies.
Various drivers for livelihood diversification have been sug-
gested, some of which are related to the environment, e.g. environ-
mental seasonality (Chen 1991) and risk of famine (Block & Webb
2001), while others are cultural, such as access to resources (Ellis
& Allison 2004) and ethnic differentiation (Smith et al. 2001), as
well as both wealth and poverty, but for different reasons (Ellis 2000;
Reardon et al. 2000). Some of these will be discussed in relation to
drivers for biological diversity in the following section.
The societal importance of cultural diversity is increasingly
acknowledged. In the UNESCO universal declaration on cultural
diversity (UNESCO 2001: 61–4), it is considered a central driver for
a number of human values: as a carrier of the common human her-
itage, as a source of creativity, as a factor in development, and as a
facilitator for social cohesion, international solidarity, and peace.3

7.4 The Concept of Biocultural Diversity


Biocultural diversity is the academic convergence of biology and cul-
ture dealing with variation and difference. The term was coined by
anthropologists, primarily working in the areas of ethnology and eth-
nobiology (Dasmann 1991; McNeely & Keeton 1995; Posey 1999).
Maffi and Woodley (2010: 5) define biocultural diversity as: ‘diver-
sity of life in all of its manifestations – biological, cultural, and
linguistic – which are interrelated (and likely co-evolved) within a
complex socio-ecological adaptive system’. This definition contains
several assumptions: (1) that humans are a part of the environment,
not apart from it; (2) that together, Homo sapiens and other life forms
3
www.unesco.org/new/en/culture/themes/endangered-languages/
biodiversity-and-linguistic-diversity/

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
8
1

188 Lennartsson et al.

have influenced one another in shifting environments; and (3) when


humans are involved, it is imperative that the past, present, and
future of these human-environment interactions are conceptualised
into a functioning whole.
Biocultural diversity is already a widely used term. Some
researchers focus on the relationship between cultural values and
biodiversity conservation, while others examine the roles of human
expression, such as language, cultural practices, and knowledge, in
using, maintaining, and enhancing biodiversity.
Biocultural diversity was originally, and is still to some extent,
related to indigenous and local societies. A need to extend the concept
of biocultural diversity beyond these has, however, been expressed
in several studies. The specificity of the term indigenous disquali-
fies a number of societies that maintain high biocultural diversity
(Cocks 2006; Pretty et al. 2009), and risks excluding mechanisms by
which cultural practices linked to the use of natural resources can be
maintained despite globalisation (Cocks 2006; Barthel, Crumley, &
Svedin 2013).

7.5 Analysis: Comparing Biological and


Cultural Diversity
By comparing biological and cultural diversity, and interpreting
each in the context of the other, both similarities and differences
are found.

7.5.1 General Differences


First, we must establish some fundamental differences between bio-
logical diversity and cultural diversity. In the long term, biological
diversification is caused by evolutionary differentiation, ultimately
leading to speciation. In the short term, diversification is caused by
succession, i.e. the gradual colonisation and establishment of spe-
cies, functional groups, etc. In both cases, biological diversity com-
prises a number of different types of organisms defined by their
genetic properties. In the short term, biological diversity can only

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
9
8
1

Diversity in Ecological and Social Contexts 189

change through colonisation, extinction or replacement of species, or


change in the relative abundance of different species. Cultural diver-
sity, in contrast, is the result of people’s thoughts and actions, such as
inventions, behaviour, and knowledge transfer and traditions, mak-
ing rapid change possible without the involvement of migration or
other demographic events.
Another, related fundamental difference is that humans, unlike
ecological systems, are able to plan ahead. In ecological systems the
processes of adaptation are affected by environmental conditions at
a given time and place. People, on the other hand, can choose long-
term security over immediate optimisation, which will be illustrated
in our examination of eighteenth-century Pennsylvanian yeomen.
This does not argue that all cultural diversity is the result of
planning. On the contrary, Ellis (1998) stresses that diversification
of livelihood may occur as an involuntary response to a crisis. This
means that the wealthy may use diversification as a strategy for accu-
mulation, while the poor may be forced to diversify to survive.
Another difference is that people can use institutional means of
distributing resources or costs among individuals in a group, which
is important for the relationship between biocultural diversity and
a secure life. There is an expectation that a society with a diverse
resource base is more resilient to fluctuations than one relying on
one or a few resources only. However, from an individual perspec-
tive, it may be better to specialise in a few resources representing the
lowest marginal risk, unless society applies institutional means of
distributing resources and ensuring security for its members.

7.5.2 Similarities and Interactions


Several of the mechanisms and effects applied to biological diversity
may provide insights also in cultural and biocultural diversity:

• The diversity of resource use may be reduced by catastrophic events,


political reforms, or other factors that break the continuity of, e.g. land-
use practices, knowledge, resource tenure, and traditions related to the
use of land and its produce (Verdery 2003; Biong Deng 2010). On the

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
0
9
1

190 Lennartsson et al.

other hand, the diversification of livelihood may be a deliberate strategy


or involuntary response to catastrophes (Block & Webb 2001; Ouma,
Obando, & Koech 2012).
• Highly variable conditions may reduce the diversity of subsistence
strategies by decreasing the benefits of marginal strategies, thus
encouraging an increased focus on fewer of the most secure strategies.
Variability may, however, also increase the diversification of resource use
in order to spread the risks. This is presumably increasingly beneficial as
security decreases (Williams 1977).
• Seasonality is one type of rather predictable variability that usually has a
negative correlation to biological diversity, but that is actually assumed
to increase the diversity of livelihood through various mechanisms.
Adaptive practices are needed to accelerate production during the limited
growth season and for storage of resources during the less productive
period. These practices also lead to diversified techniques for cultivation,
transhumance, harvesting and storing, institutions for labour exchange,
resource allocation, property rights, and social networks, as well as norms
and beliefs related to important activities, potentials, and constraints.
Furthermore, seasonality may create a need for livelihood strategies other
than those relying on ecosystem resources (Ellis & Allison 2004; Orr,
Blessings, & Saiti-Chitsonga 2009).
• Increased competition over resources, caused by, e.g. population growth
or resource depletion, can create innovative uses of resources and land
use, adoption of new technology, knowledge, and traditions, which lead to
increased biocultural diversity.
• Heterogeneity of landscapes and ecosystems can be expected to lead to
diversification of resources and land-use practices, as well as various other
cultural expressions related to resource usage.
• Conversely, diversity of traditions (e.g. land-use practices, land tenure,
product use) strongly influences and enhances biological diversity in
cultural landscapes (Plieninger, Höchtl, & Spek 2006). The example
of the Romanian village discussed later in this chapter demonstrates
how diversity of land-use traditions creates biocultural diversity in the
surrounding landscape. Maurer and colleagues (2006) find that biocultural
diversity differs between Swiss landscapes, depending on which of the
country’s three cultural traditions (German, French, Italian) dominate
land-use practices. In Central and Eastern European countries, the choice

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
1
9

Diversity in Ecological and Social Contexts 191

of post-communist land-reform strategy is an important determinant of


biocultural diversity. The choice of strategy depends on environmental
conditions, but is also strongly influenced by various cultural traditions
(Swinnen 1999).

7.5.3 Differing Perspectives


Research on cultural diversity frequently includes approaches from
ecology. Ecological anthropology has used ecology and evolutionary
theory as analogies for discussing the evolution of cultural diversity (e.g.
Jochim 1981: ch. 1), and the importance of the environment in general
is acknowledged. In research on biological diversity, in contrast, human
influence is often absent. Most of the ecological models discussed pre-
viously in this chapter, which attempt to explain the cause or effect of
biological diversity, overlook the human influence on biological sys-
tems. It is clear that land use or other forms of human influence on
nature would have an impact on most of the models, indicating that
biological diversity is strongly determined by several aspects of biocul-
tural diversity, especially land use (e.g. Eriksson 2013). The absence of
human impact in biological-diversity models constitutes a problem in
ecological theory, particularly when dealing with the restoration and
conservation of biodiversity in landscapes with a long history of human
impact. This aspect is further discussed in Chapters 5 and 9.
The scarcity of ecological theory incorporating both natural
and anthropogenic processes is primarily the result of ecologists’
desire to simplify their models. However, when these models are put
into practice, a certain intentionality can sometimes be observed.
One such example, often discussed in restoration ecology, is the con-
cept of rewilding, i.e. to reintroduce or imitate pre-cultural ecological
conditions (Soulé & Noss 1998). For rewilders, the reintroduction
of large mammal herbivores and predators (‘vertebrate megafauna’)
and of ‘natural’ processes is a goal in itself, with no defined manage-
ment outcome (Monbiot 2013). Therefore, because of the devastating
impact of humans on the desired megafauna, ‘rewilding is essentially

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
2
9
1

192 Lennartsson et al.

about mitigating anthropogenic ecosystem impacts’ (Sandom et al.


2013; Jörgensen 2015). Usually, however, biological conservation
practices show a great awareness of the significance of biocultural
diversity, although rarely expressed through this particular term.
Among practitioners of biodiversity conservation or cultural heri-
tage conservation, human-nature links have long been acknowledged
and are considered essential for successful conservation (e.g. Beaufoy,
Baldock, & Clark 1994; Ikonen & Lammi 2000; Oppermann, Beaufoy, &
Jones 2012). In the international conservation arena, the necessity
of linking nature with culture is, moreover, explicitly expressed (e.g.
UNEP 2007; IUCN 2008: 49).
Despite an awareness of the importance of traditional land use
for biodiversity among conservation practitioners, there is, as dis-
cussed in Chapter 5, a risk that the lack of ecological theory on land
use leads to a general portrayal of biodiversity as ‘best’ in ‘pristine’
ecosystems (e.g. Kricher 2009; Leadley et al. 2014: ch. 5). Indeed land
use, including many forms of traditional land use, may not favour
biodiversity. Yet, if a certain type of traditional land use has a long
history, biodiversity dependent on such land use can be expected to
have developed (see further Chapters 6 and 9).

7.5.4 Understanding the Mechanisms of Human-Nature


Relationships
The convergence of cultural and environmental diversity can be seen
in many different contexts and on many scales, including land-use
and ecosystem diversity (e.g. the Romanian example in this chap-
ter) and linguistic and resource-use diversity (Smith 2001; Skutnabb-
Kangas et al. 2003), as well as geographical correlations between high
ethnic diversity and biodiversity (Pretty et al. 2009).
One important task for the study of biocultural diversity is
to understand the drivers for the emergence and loss of biocultural
diversity, as well as the significance of it. In order to fully understand
these two issues, knowledge of the causal mechanisms of human-
nature relationships is required. Researchers may therefore need to

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
3
9
1

Diversity in Ecological and Social Contexts 193

focus on local land-using communities where such reciprocal effects


of environment and culture are most common and easy to detect (see
Chapter 4). This would provide an opportunity to examine diversity
in how local land use and ecosystems have developed simultaneously
and how they are linked through dynamic feedback mechanisms.
Landscapes and resources have been shaped by land use along with
the knowledge of how to use the resources.
This chapter will now investigate causal relationships between
cultural and biological diversity through four case studies. The first
examines the contemporary diversity of ecosystem resources in a
highly seasonal and dynamic Scandinavian environment used by
reindeer herders. The second relates historical crop and livestock
diversification in North America to production security and labour
needs. The third discusses the importance of institutional diversity
for the use of natural resources in a Swedish historical transhumance
system. Finally, the fourth case study investigates biocultural diver-
sity in a contemporary Romanian mountainous agrarian society,
which exemplifies mechanisms for the reciprocal influences between
cultural and biological diversity.

7.6 Four Examples of Biological


and Cultural Diversity
7.6.1 Reindeer Husbandry in Sweden
Reindeer husbandry is a traditional form of land use and meat pro-
duction in northern Sweden. In line with the Reindeer Husbandry
Act, this is carried out within fifty-one herding districts, of which
the majority stretch from the Norwegian border in the west to the
coast of the Baltic Sea in the east. These herding districts are geo-
graphical units, where reindeer herders have collective grazing rights
(they do not own the land). The districts are also legal and economic
entities. Since 2006, these herding districts have been organised
under the Swedish Sami Parliament. In most districts, a migratory
herding system is used where semi-domesticated reindeer graze the

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
4
9
1

194 Lennartsson et al.

Figure 7.1 Reindeer summer pastures at Bietsávrre settlement,


Lappland, Sweden.
Photo: Tommy Lennartsson.

alpine vegetation of the western mountains during the summer


(Figure 7.1), and winters are spent on the lichen-rich pine heaths
of the boreal forests further east (Figure 7.2; Danell 2000; Moen &
Danell 2003; Moen 2008; Kivinen et al. 2010; Moen & Keskitalo
2010). Generally, grazing conditions related to weather and forage
availability, together with land access and disturbances from other
land uses, are prime concerns for sustainable reindeer husbandry
(Moen & Keskitalo 2010).

7.6.1.1 Winter Forage and Forestry


The limiting key resource for reindeer husbandry is the availability
of lichens during the winter: ground lichen which the reindeer find
under the snow and, during periods with difficult snow conditions,
arboreal lichen. The animals are adapted to use lichens as a source of
energy to ensure survival during the winter. Body growth only occurs

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
5
9
1

Diversity in Ecological and Social Contexts 195

Figure 7.2 A reindeer herd on winter pastures in the boreal forest,


Sweden.
Photograph Anders Esselin.

during the summer when more high-quality forage is available. Snow


conditions are crucial for the availability of ground lichen, and late-
season ice-crust formation can have a strong effect on the condition
and survival of the animals (Heggberget, Gaare, & Ball 2002; Moen,
Andersen, & Illius 2006; Roturier & Roué 2009). This has sometimes
led to the collapse of herds, in turn causing herders to leave reindeer
husbandry for other livelihoods (Brännlund & Axelsson 2011). Today,
reindeer herds are given supplemental food when grazing conditions
are poor. This is, however, associated with substantial economic cost
and is not sustainable over longer periods.
Apart from weather fluctuations, the largest effects on the
amount and availability of lichens come from large-scale commer-
cial forestry. Forestry is an important industry in Sweden, with an
export value of ca. 12 per cent of all exported goods. In northern

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
6
9
1

196 Lennartsson et al.

Sweden, 50 per cent of forested areas is owned by forest companies


(both private and state-owned), 44 per cent by private non-industrial
owners, and the remainder by municipalities or the Swedish Church.
Commercial forestry is conducted across the entire reindeer winter
grazing area, with significant effects on the landscape (Kivinen et al.
2010; Moen & Keskitalo 2010). Today, forestry is conducted through
clearcuttings with subsequent planting or natural regeneration. In
the late nineteenth century, forestry was mainly conducted through
selective cutting of the largest trees, which opened up the forest but
had a negative effect on forest regeneration (Östlund, Zachrisson, &
Axelsson 1997). To a certain extent, clearcutting was already part of
forestry strategies, but it became more common as the paper-pulp
industry developed, and has been the main forestry strategy since the
1950s (Lundmark, Josefsson, & Östlund 2013). Forestry has changed
the boreal landscape, with an increasing proportion of even-aged
young stands and few old-growth forests. Clearcutting obviously
removes habitats for arboreal lichens, and logging residue makes it
difficult for the reindeer to reach the ground lichens under the snow.
Site preparation, especially soil scarification in order to improve the
establishment of seeds or planted saplings, removes ground lichens.
Young forests are usually very dense, and the lack of light restricts
the growth of ground lichens. Forest fertilisation during the stand
rotation period is also detrimental to ground lichens since it pro-
motes other competing field-layer vegetation and the shade provided
by the tree canopy.

7.6.1.2 Mismatch of Scales


Forestry is practised on a stand scale, where a stand is a relatively
homogenous, often even-aged, part of the forest. The size of each
stand ranges between <10 hectares and a few hundred hectares. The
forest companies plan their annual harvest based on the character-
istics of the available stands, where stands of a similar age are more
or less interchangeable when planning for management activities.
The ownership patterns of Swedish forests are highly fragmented and

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
7
9
1

Diversity in Ecological and Social Contexts 197

neighbouring stands may therefore be owned by different companies


or private landowners.
In contrast, reindeer herders within a herding district are
divided into ‘winter groups’ (siidas) consisting of one or a few fami-
lies. These groups use areas on the scale of several tens of thousands
of hectares during the winter, moving the reindeer in accordance
with grazing conditions. Such winter-grazing areas have many dif-
ferent owners and they are also affected by other land uses, such as
mining, tourism, energy production, agriculture, etc.
This mismatch of spatial scales between the forestry industry
and reindeer herders makes planning and resource management dif-
ficult, especially for the reindeer herders. By law and/or sustainable
forest management regulation (e.g. Forest Stewardship Council), for-
estry companies are obliged to consult with reindeer herders before
they begin harvesting. There is no requirement, however, for the two
parties to agree on harvest plans. Herders therefore perceive these
consultations as fora for information where they can influence a few
details, but not overall harvest levels, etc. (Sandström et al. 2006).
Furthermore, since landownership is so fragmented, the herders
of a particular district may need to attend consultations with sev-
eral companies, making the landscape perspective of the siida land
use difficult to uphold. In recent times, most herding districts have
developed GIS-based reindeer management plans using their own,
local knowledge. These plans make it easier to describe the herders’
dynamic use of the landscape in their consultations with the forest
companies. Yet this seems to have only marginally improved the out-
come of consultations for the herders as the most critical aspects,
such as overall harvest levels, are not part of the discussions.

7.6.1.3 Temporal Dynamics, Spatial Variation, and


Diversity of Resources
A Sami herder’s pasture is a dynamic entity, for example due to vary-
ing snow conditions during or between winters. Forest stand struc-
ture, snow quality, and forage accessibility are considered when

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
8
9
1

198 Lennartsson et al.

herders evaluate foraging possibilities during the winter. It is often


assumed that clearcuts will have a thicker and denser snow cover
than forest stands, as the forest canopy will capture some of the snow
(e.g. Helle, Aspi, & Kilpelä 1990). The Sami approach is, however,
much more dynamic and allows for spatial and temporal variabil-
ity. For instance, the timing of warmer periods during the winter in
relation to snow accumulation in the forest canopy will determine
if forest stands have a higher or lower risk of ice-crust formation. If
there is still snow in the canopy, it will drop at some stage, result-
ing in hard-packed patches close to the tree trunks (Roturier & Roué
2009; Horstkotte & Roturier 2013). On the other hand, if the wind
has already removed the snow from the canopy, warm spells will
have a stronger effect on ice-crust formation within clearcuts.
The diversity of the landscape is crucial for reindeer husbandry
as it allows for rotational grazing in accordance with the dynamics
of the lichen resource and weather variation. This is also an impor-
tant aspect of the resilience of the natural resource use of the herd-
ers (Moen & Keskitalo 2010). Variations in forest-stand structure,
together with variations in wind and temperature, will create diverse
grazing conditions, constantly shifting within and between years.
Sustainable reindeer husbandry is highly dependent on a maintained
landscape diversity and on a diversity of management tools and deci-
sions based on rich traditional ecological knowledge.

7.6.1.4 Knowledge of Diversified Land Use


Reindeer herders’ traditional ecological knowledge of ‘pastures’
clearly exemplifies how important it is to understand and use land-
scape diversity. When referring to pastures or grazing, Sami herders
often use the word guohtun (Roturier & Roué 2009), which includes
a great deal of information, not only amount of forage, but also acces-
sibility and reindeer behaviour in relation to the forage. According to
Roturier and Roué (2009), herders have given the following definition:

Guohtun is more than just the food, it is the whole actually: how
the reindeer can get their food. It means that what is above the

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
9
1

Diversity in Ecological and Social Contexts 199

lichen [i.e. the trees and the field layer vegetation] is also included
in it, everything that influences the food itself . . . Guohtun means,
in fact, that the reindeer get the food. Therefore it includes the
snow conditions, snow depth, the ground, what is on the ground, if
there is lichen: the idea is that reindeer can get their food: guohtot.
The reindeer guohtot: the reindeer eat, the reindeer graze.

Horstkotte, Roturier, and Moen (2014) define landscape diversity as


the various structural and functional attributes that exist in space
and time and across a number of scales. They further divide land-
scape diversity into three dimensions: environment, society, and
governance. Future policies on this multidimensional landscape
may have two different effects on reindeer husbandry: (i) main-
tained landscape diversity will enable herders to react to distur-
bances by upholding or enhancing mobility in the landscape (as
discussed in relation to the guohtun), or (ii) a fragmented and less
connected landscape will reduce the capacity of reindeer herding to
adapt to disturbances and other external factors by restricting herder
options. In other words, different types of diversity can increase or
decrease adaptation to disturbances. Flexibility of this kind is as
important today as it was in the nineteenth century (Brännlund &
Axelsson 2011).

7.6.2 Managing Biocultural Diversity in Colonial America


The yeomen of eighteenth-century Pennsylvania inhabited the fron-
tier of the British Empire. Like any colony, their purpose was to pro-
vide food and raw materials for the international market in order to
create tax revenue for the Crown. For the colonists, it was an oppor-
tunity to gain affordable land. In Chester County, Pennsylvania,
these new landowners included the descendants of English small-
holder yeomen families, a few gentlemen of means, Welsh sheep-
herders, urban craftsmen from London, husbandmen – highly trained
labourers – from central and southern England, and ‘Dutch’ labourers
and craftsmen from what is today southwestern Germany. Despite
their diverse origins, these yeomen practised a common form of

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
0
2

200 Lennartsson et al.

Figure 7.3 An Indian summer view of the farm and stock of James
C. Cornell, 1848, by Edward Hicks. A black-and-white version of this
figure will appear in some formats. For the colour version, please refer
to the plate section.

agriculture which merged the Native American practice of shifting-


field horticulture with the growing of European market crops.
Wheat was grown for the markets in London, Caribbean planta-
tions, and the cities of southern Europe, while rye, buckwheat, and
barley were consumed locally. Oats, spelt, and Indian corn (maize)
were grown in small quantities to feed horses and fatten swine. Most
cattle and sheep were left to graze in fallow fields and browse in the
woods (Figure 7.3).

7.6.2.1 Labour in Colonial Pennsylvania


Pennsylvania yeomen sought to balance local needs with the need
for market profits. Land was plentiful and relatively inexpensive, but
labour was scarce and came at a relatively high price. One of the
reasons for the high labour costs was the ready availability of land.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
1
0
2

Diversity in Ecological and Social Contexts 201

Except for slaves and bound servants, a labourer worked for others
only as long as it took to establish a sufficient reputation to convince
a local gentleman to back a mortgage for a piece of land. As a result,
the uncommitted labour pool consisted mostly of teenagers and very
recent immigrants.
The costs of training field hands and maidservants in colonial
Pennsylvania are reflected in the valuations of servants contained in
contemporary probate inventories. When a propertied person died, a
list of their assets was compiled by neighbours appointed by a local
magistrate. The time a servant had left to serve was listed as an asset
of the estate. Servants were typically contracted for a number of years.
Seven-year agreements were common. Maidservants tended to learn
their tasks more quickly than field hands, and were therefore given a
higher valuation during their initial years of service. The dairy market
in colonial America was undeveloped compared with the grain mar-
ket, and the greater potential proceeds of the half-trained field hands
made them more valuable than fully trained maids. The physical
demands and complexity of working diverse crops meant that a field
hand reached maximum proficiency only in the last year of his service.

7.6.2.2 Crop and Livestock Diversity


One way of reducing training time and being able to employ labour-
ers of all skill levels is by reducing crop and livestock diversity.
Mono-cropping and specialised herds have the lowest potential costs
per unit because they make it easier to use lower-skilled labour,
gain efficiency through economies of scale, simplify scheduling of
work, and chase the highest market price. However, this strategy also
comes with higher risks. A lack of diversity leaves crops susceptible
to pests and incompatible weather, as well as human-related hazards
such as labour shortages at peak harvest and rapid market fluctua-
tions. Crop diversity provides greater flexibility in work schedules,
spreads labour requirements, provides a balance between sustenance
and market crops, and broadens the range of options for managing
changing conditions.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
2
0

202 Lennartsson et al.

Figure 7.4 Comparison of acreage of wheat, meslin (mixed wheat and


rye), and rye, Chester County, USA.
Source: probate inventories, 1713–89.

A signature method of crop diversity practised by Pennsylvania


yeomen was the inter-cropping of wheat and rye in the same field, a
widely adopted practice in the mid-eighteenth century (Figure 7.4).
By sowing these winter grains in the same field, the yeomen reduced
the total potential production of marketable wheat, but insured parts
of the crop from cold and dry springs, which can damage wheat. The
hardy and fast-sprouting rye sheltered the wheat until more favour-
able weather arrived. The resulting rye had a value on the local mar-
ket, and in most homes it was mixed with coarse wheat ‘middlings’
for a hearty brown bread. Since antiquity, bread made of mixed grains
has been known as ‘meslin’, and agricultural historians have adopted
the word to also describe fields of mixed grain.
Initially, Pennsylvania yeomen attempted to grow as much
wheat as possible, and between 1713 and 1720 it composed more than
80 per cent of their crops (Table 7.1). After 1730, Chester County yeo-
men initiated a more sustainable strategy by growing more rye and,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
3
0
2

Table 7.1 Mean Acreage of Grain Per Grain Producer in Chester County, Pennsylvania According to Probate
Inventories and Diversity (Shannon Index)

Period Mean acreage (%) Crop diversity

Wheat Meslin Rye Indian Corn Barley Oats Buckwheat

1713–19 16.8 (81) 0.9 (05) 0.2 (01) 0.7 (03) 0.5 (02) 1.2 (06) 0.4 (01) 0.8

1720–9 14.2 (88) 0.8 (05) 0.4 (02) 0.0 (00) 0.2 (01) 0.5 (03) 0.0 (00) 0.4

1730–9 14.3 (73) 2.8 (14) 0.7 (03) 0.8 (04) 0.2 (01) 0.9 (05) 0.0 (00) 1.0

1740–9 13.5 (60) 6.2 (28) 1.1 (05) 0.8 (04) 0.1 (00) 0.7 (03) 0.0 (00) 1.1

1750–9 15.0 (60) 5.6 (22) 1.3 (05) 2.4 (10) 0.1 (00) 0.6 (02) 0.1 (00) 1.4

1760–9 10.6 (46) 7.0 (31) 2.1 (09) 2.2 (10) 0.1 (00) 0.7 (03) 0.2 (01) 1.3

1770–9 10.6 (50) 4.6 (22) 2.7 (13) 2.0 (10) 0.1 (00) 0.9 (04) 0.3 (01) 1.4

1780–9 9.6 (34) 3.6 (13) 3.0 (11) 9.6 (34) 0.1 (00) 2.2 (08) 0.4 (01) 1.4

1713–89 15.2 (63) 3.9 (16) 1.4 (06) 2.3 (10) 0.2 (01) 1.0 (04) 0.3 (01)

terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007


N.B. italics designate estimations from sparse data.

203

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
4
0
2

204 Lennartsson et al.

as animal fodder, Indian corn and oats (Table 7.1; Figure 7.4). The
non-logarithmic Shannon Index of yeoman’s crops in probate inven-
tories in Table 7.1 confirms a diversification of crops. The depressed
market for wheat (Figure 7.4) may be sufficient explanation for the
diversification of grains, but although prices rose throughout the
second half of the century, Chester County yeomen did not return
to wheat. There were likely many other reasons for these changes,
involving responses to changing regional climate, social agendas,
and land availability (Scholl 2008; Scholl, Murray, & Crumley 2010).
Yeomen continued to profit from the wheat market in the second
half of the eighteenth century, but individually grew less and less
of it. They made adjustments to their quantities of local and animal
crops, but maintained a consistent level of diversity crop between
1750 and 1789.
While thus the Pennsylvania yeomen would readily change
their cropping patterns, they kept the sizes of their herds fairly con-
stant throughout the century (Table 7.2; Figure 7.3), despite signifi-
cant market pressures. This meant that herds were carefully culled
in the early part of the century to keep numbers under control, and
then, as the sizes of holdings were reduced in the latter half of the
century, precious labour was devoted to increased fodder production.
Historical examples like this one are useful because they
occurred at an observable scale. Thousands of households in one cen-
tury of a county’s history can be more easily analysed from the distance
of time than the present hundreds of thousands of households which
form the current global agricultural market. Eighteenth-century
Pennsylvania may not show us that modern industrial agriculture
should mix its grain, but rather that agrarians are embedded in social
and environmental contexts limiting their choices and defining their
goals. If we value agricultural biocultural diversity, we need to pro-
vide agriculturalists with local control. At the scale of a country, the
world market can provide enormous diversity and savings by sup-
plying relatively low-cost goods from other nations. However, these
cheap goods are often the result of short-term production strategies

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
5
0
2

Table 7.2 Mean Livestock Herds and Fodder in Chester County, Pennsylvania Inventories and Diversity
(Shannon Index), 1713–1789

Period Mean number of livestock Number of inventories Livestock diversity

Horses Cattle Swine Sheep With Stock With Hay (%)

1713–19 4.3 6.4 9.2 12.7 63 2 (03) 1.3

1720–9 4.8 8.6 7.3 18.4 188 15 (08) 1.3

1730–9 5.1 9.7 6.6 20.2 245 33 (13) 1.2

1740–9 4.4 9.4 7.1 15.1 458 121 (26) 1.3

1750–9 4.5 8.1 7.5 14.4 378 122 (32) 1.3

1760–9 3.7 8.5 7.8 15.3 498 183 (37) 1.3

1770–9 3.6 7.4 7.1 13.6 484 175 (36) 1.3

1780–9 3.9 7.3 7.6 11.1 488 198 (41) 1.3

terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007


1713–1789 3.9 7.9 7.1 13.7 2802 849 (30)

205

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
6
0
2

206 Lennartsson et al.

which include exploitation of labour and land, and undercut the cost
of local production. The likely result is a reliance on food produced in
distant places under conditions of higher than necessary risks.

7.6.3 Institutional Diversity in Swedish


Summer Farming ca. 1600–1900
Institutions provide a structure for most human interaction and its
use of natural resources. A broad definition offered by Elinor Ostrom
(2005: 3) states that ‘institutions are the prescriptions that humans
use to organize all forms of repetitive and structured interactions.’
Douglass C. North (1991: 97) sees institutions as ‘humanly devised
constraints that structure political, economic and social interaction’
which ‘consist of both informal constraints (sanctions, taboos, cos-
tumes, traditions, and code of conduct), and formal rules (constitu-
tions, laws, property rights)’. In other words, institutions represent
sets of rules used by individuals to organise recurring activities,
which in many cases include interaction with nature. A clear under-
standing of institutions is therefore crucial for the study of historical
ecology.
Sustainable land management requires institutions that are
adapted both to the users and to the resources. The complex interac-
tion between humans and nature and the multifaceted grid of institu-
tions that shapes it are, however, difficult to understand. Users are
often unaware of the rules and norms they adhere to. To complicate
matters, institutions responsible for natural resource management
exist on multiple levels and scales, where day-to-day decisions are
found on one level, while fundamental rules governing who is eligi-
ble to withdraw resources are on another (Ostrom 2005: 58–62).
If understanding the roles of these institutions in contempo-
rary society is hard, understanding their roles in the past is even more
challenging, as oral communication with users from the past is not
possible. Historical ecology, however, provides an opportunity to
examine long-term change in institutional settings and its influence
on the use of natural resources and impact on biocultural diversity.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
7
0
2

Diversity in Ecological and Social Contexts 207

This approach will not only shed light on how institutional diversity
of the past is linked to natural resource management, it will also help
us understand present-day resource use, as well as provide guidelines
for future usage of natural resources (de Moor 2015).
Institutions are present in all natural resource management
regardless of ownership system: private, public, common, or a mix
of these. In the past decades, management of common-pool resources
(CPRs) has, however, been the focus of intense discussion involving
institutions, diversity, and sustainability. There are two main rea-
sons: Firstly, since the 1980s, research has changed the perception of
how people manage resources held as commons and work together.
Until the 1970s, the predominant view of commons was that they
trapped people in social dilemmas which would deplete CPRs.
Individuals would therefore ‘free ride’ and extract more resources
than an area could produce. This view, labelled ‘the tragedy of the
commons’ after an article by Garrett Hardin (1968), can be traced
back as far as Aristotle. Yet today we know that humans, by self-
governing, can manage CPRs whilst sustaining ecological systems
(Ostrom 1990). Institutions make it possible for user groups to define
social interaction and competition. They make the behaviour of
other users more predictable and reduce risk and uncertainty (North
1990). Institutions have not always been successful in managing
common-pool resources, but neither have private and state owners
(Feeny et al. 1990; Dietz, Ostrom, & Stern 2003). The second rea-
son for the focus on common-pool resources is our current awareness
that many of the most crucial resources on earth are held in com-
mon, such as air, oceans, and watersheds, and ultimately the entire
planet. Even if these larger commons differ from the CPRs described
earlier in this chapter, most importantly in the sense of excludability,
the knowledge we can gain from studying small and medium-sized
CPRs may enhance humanity’s management of these larger com-
mons (Stern 2011).
Numerous examples from around the world prove that it is pos-
sible for user groups to self-govern resources in a sustainable manner

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
8
0
2

208 Lennartsson et al.

(e.g. Netting 1981; Siy 1982; McKean 1986; Ostrom 1992). Results
from field studies are supported by experimental studies and game
theory (Ostrom, Gardner, & Walker 1994). Together these efforts
have led to the development of comprehensive frameworks, theo-
ries, and methods dealing with questions regarding CPRs and coop-
eration (Ostrom 1990, 2005). Major conclusions are that a diversity
of institutions is needed and that institutions that manage various
natural resources need to be able to adapt to local conditions, natural
as well as social and cultural. This means that no single solution can
be found. In what follows, we will examine long-time institutional
changes by analysing how a transhumance system which used forest
pastures was adapted to new local conditions and changes in national
legislation.

7.6.3.1 Swedish Summer Farms


For an economic system to be viable, its management institutions
must be able to adapt to new conditions. A lack of adaptability in
the institutional settings could damage or even deplete the resources
used. Here we will study a transhumance system from the sixteenth
century to the beginning of the twentieth century and examine
how it adapted to new conditions, as well as its final failure and
disappearance.
A transhumance system based on what we here call summer
farms (shielings), emerged in Sweden in the sixteenth and seven-
teenth centuries, after the late medieval crises (Larsson 2009).
A summer farm was a summer settlement for the purpose of grazing
common pastures away from the home village, and where milk was
processed into long-lasting products (e.g. butter, cheese, and whey
cheese). There were dwelling houses, byres for the livestock, and spe-
cial buildings for milk processing (Figure 7.5). The Swedish summer
farm system is classified as an alpine transhumance where animals
were commonly stabled during winter and the collection of winter
fodder for the livestock was an essential summer activity (Davies
1941; Larsson 2012). The system shows similarities with other

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
9
0
2

Diversity in Ecological and Social Contexts 209

Figure 7.5 A Swedish summer farm community consisted of members


from many households using the same settlement and area. Herders,
cows, and goats at Björnbergets summer farm in the parish of Leksand,
Sweden, early twentieth century.
Photo: Hans Per Persson, Djura, Leksand. Courtesy Lars Liss Photo
Archive in Gagnef, Sweden.

regional alpine transhumance systems around the world, although


the Swedish version was mostly found in boreal regions and ecosys-
tems. A characteristic regional feature of Swedish summer farming
was the use of female herders (Figure 7.6; Larsson 2014b). From a
biocultural diversity perspective, Swedish summer farms comprised
very diverse ecosystems, both regarding biodiversity and use, e.g. the
systems for grazing and the collection of mire hay, leaf, lichen, etc.
for winter fodder. From the seventeenth to the twentieth century, the
management of the summer farms was mostly based on cooperation.
Peasants not only shared the pastures, they also cooperated when
tending the animals in the forest pastures and when processing milk.
The successful management of the commons within a summer
farm system needed rules, i.e. institutions (see also Chapter 4). These

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
0
1
2

210 Lennartsson et al.

Figure 7.6 Detailed rules for where and when the grazing cows could
be herded were necessary institutions for a sustainable use of woodland
and mire pastures. Female herders following the herd from Björnbergets
summer farm in the parish of Leksand, Sweden, early twentieth
century.
Photo: Hans Per Persson, Djura, Leksand. Courtesy Lars Liss Photo
Archive in Gagnef, Sweden.

could be rather simple when the users were few, but had to be devel-
oped when the number of users increased. Rules restricting usage
ranged from written laws passed by the diet to formal and informal
rules developed by the users. Court records provide an insight into
the practical application of laws, although informal rules are harder
to trace as they were never written down.
The overall goal for the farmers was to ensure a balance
between appropriation and provision, in other words, to protect
their land from overuse. Boundaries for excluding others were
an important way of protecting resources, and two types of such
boundaries have been identified: (1) the physical boundary of
the resource, and (2) the rules governing individual access rights
(Ostrom 1990, 2005).

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
1
2

Diversity in Ecological and Social Contexts 211

7.6.3.2 Defining the Resource Area


When farmers first established summer farms in the vast boreal for-
ests of central Sweden, there was an abundance of forests and mires
available for grazing. From the sixteenth century until the middle
of the seventeenth century there were no clear boundaries between
the different summer farms; the few boundaries mentioned before
1680 concerned much larger areas. Not even parish boundaries were
related to grazing rights. When the population and the number of
summer farm users increased, the farmers acted to protect their com-
mons, mainly by producing detailed definitions of their territory.
After 1680, the establishment of resource areas for each user group
became important, and the nature of the commons thus changed from
open access to a common-pool regime (Bromley 1991). The establish-
ment of resource areas for the summer farms was in an intense phase
until ca. 1730. This process was initiated by the users, but supported
and upheld by local courts, where boundary disputes were settled
(Larsson 2014a).
At the end of the seventeenth century, a major change in the
use of the commons occurred, which further increased the need for
clearly defined boundaries. In areas with high population density,
all farmers were required to have a summer farm. Allowing some
peasants to stay in the village with livestock while others moved
their livestock to summer farms would have posed a threat to the
agricultural system. It would have meant higher costs for the sum-
mer farm users, and also the impending risk that village pastures
would be depleted when the summer farm users returned in the
autumn. The pastures were to be shared by all users, and by law
all landowners had the right to use the commons. As the summer
farms became compulsory, however, tension grew between old and
new summer farm users, and unresolved conflicts were brought to
the local courts. In the first decades of the eighteenth century, the
summer farms’ boundaries and individual access to these farms
were more clearly defined. With the increasing commons usage in
the second half of the eighteenth century, further improvements

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
2
1

212 Lennartsson et al.

were required. Various strategies of protecting the farm boundaries


were developed, e.g. making them more visible by building fences
and imposing fines for trespassing. Well-established boundaries and
well-defined user groups, however, proved insufficient protection,
as the user groups themselves also posed a threat to the resources
of the summer farms.

7.6.3.3 Restrictions of Use, Adapted to Local Conditions


The appropriation rules continuously became stricter, from the time
summer farms were introduced until their rapid decline in the late
nineteenth century. This included continuous fine-tuning of the reg-
ulations in order to match changing local conditions. For example,
one way of protecting the pastures of a transhumance system is to
restrict the length of the grazing season by regulating the dates when
livestock should be moved from the villages in the spring and also
for its return in the autumn. These rules were increasingly enforced
around 1700 when it became compulsory for peasants to have a sum-
mer farm. Over time, regulations were gradually more adapted to
local conditions. The right to decide which date users had to move
their livestock was transferred from the villages to the individual
summer farm communities. Other ways of adapting were to impose
limits on the number of livestock each member could have and to
regulate how many hours each farmer needed to spend improving the
pastures each year. From the second half of the eighteenth century,
summer farm communities started writing down their own local
rules, which in practice functioned as by-laws.

7.6.3.3 Regulating How to Make Rules


Rules in themselves are not enough. The process of establishing rules
is highly important for people’s willingness to comply with them.
‘Collective-choice arrangements’, which allow users to influence the
rules, are needed (Ostrom 1990, 2005). When the number of sum-
mer farms increased during the seventeenth century, a new institu-
tion was introduced to manage them – the summer farm community.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
3
1
2

Diversity in Ecological and Social Contexts 213

Large summer farms, consisting of up to forty households, were


sometimes subdivided, often into quarters, and some decisions about
farm management were delegated to these groups, as long as deci-
sions did not affect others. Rules that could have impacts for other
user groups, moreover, had to be confirmed in the local courts. New
summer farm regulations were announced in the parish church so
that people could contest them. Other rules concerned both summer
farms and villages and had to be introduced through other types of
fora, such as the parish councils (Sw. sockenstämma).

7.6.3.4 Monitoring, Trial, and Sanctions


Despite rules being approved and established through consent, some
summer farm users may have been tempted to break them. Regardless
of the reasons for such violations, a user group needed to monitor
its own members, as well as to make sure that others respected the
boundaries of the resource area. Monitors actively had to assess the
condition of the resource and the users’ behaviour, in order to be suc-
cessful (Ostrom 1990, 2005). From the seventeenth century summer
farm users monitored their own resources closely. This work was
only a small extension of the regular work of tending the grazing
livestock and therefore came at a low cost. As the use of the pastures
increased, it became more important to prevent members of a sum-
mer farm community from overharvesting and also from neglecting
their duties. The communities appointed monitors or a warden who
were rewarded by being allowed to keep a share of the fines collected.
The summer farm communities gradually gained more and more
power to collect fines from their members. The by-laws also became
more detailed and the number of potential violations therefore also
increased.
In times of conflict within or between summer farm communi-
ties, an institution for resolving such disputes was necessary. Most
disagreements could be resolved through informal meetings within
the summer farm community, but some were brought to the local
courts (Sw. Häradsrätten). These courts formed the lowest tier of

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
4
1
2

214 Lennartsson et al.

the Swedish judicial system and were led by judges, assisted by lay
judges chosen from among the local farmers. These lay judges there-
fore had remarkable insight into local customs and conditions. The
local courts became low-cost and successful arenas for solving prob-
lems arising from the management of common pastures. Their suc-
cess was based on the interplay between de jure rights and de facto
rights, i.e. rights given lawful recognition and rights created by users
to facilitate resource management (Larsson 2014a). During the nine-
teenth century, the courts’ role in resolving conflicts between sum-
mer farm users diminished, mainly because the courts started paying
less attention to de facto rights in their verdicts.

7.6.3.5 The End of Summer Farming


The summer farm users showed a remarkable ability to cope with
changing conditions from the sixteenth until the end of the nine-
teenth century. The diversity of institutions made it possible to
adapt to new conditions and the interaction between the users and
the local government was a key to success. Yet the evolved summer
farm system of the eighteenth and nineteenth centuries came to an
end in the first decades of the twentieth century. The main reason
for its rapid demise was a new agricultural system, e.g. the introduc-
tion of crop rotation, mechanisation, changing patterns in the con-
sumption of dairy products, as well as land reforms through which
common grazing areas were transferred into private ownership.
Changes outside the agricultural sector also had a negative impact
on the summer farms. One example is the new markets for timber,
which changed the focus on forest products from pasture to timber,
and resulted in intense campaigning and propaganda against forest
grazing. Forests changed from being an intrinsic part of agriculture
to being a natural resource that needed protection from agriculture.
A second example is the change brought by the textile and fash-
ion industries after 1850; as cotton became the staple raw material,
the demand for sheep and goat products fell rapidly (Blomkvist &
Larsson 2013).

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
5
1
2

Diversity in Ecological and Social Contexts 215

7.6.4 Ecosystem Diversity and Land-Use Diversity in


Romanian Grassland-Based Agriculture
In the Romanian Carpathians, one of the largest areas in Europe char-
acterised by agriculture independent of fossil fertilisers and fuels is
found. Here, production relies on low-input semi-natural grasslands
as a base for animal feed, and on manure, which provides the nutri-
ents for arable fields. The agricultural landscape, for example in the
village of Botiza in the northern region of Maramureș, shows diver-
sity of different kinds and scales.

7.6.4.1 Types and Scales of Diversity


On the largest macro scale, the village exhibits a variety of general
land-use types over gradients in the landscape. Close to the village
centre there are small garden plantations, followed by areas of culti-
vated fields. While most of the landscape around the village consists
of hay meadows, further away, in the mountains, there are the sum-
mer pastures for cows, sheep, and horses (Figure 7.7).
Each of these landscape zones contains a great diversity of
anthropogenic ecosystems forming an intricate patchwork. The
arable fields are used for various crops (maize, beans, potatoes, cere-
als) and are in different stages of a rotation cycle involving fodder
crops such as clover, alfalfa, and semi-natural vegetation. The diver-
sity of the hay meadows is visible already from a distance as differ-
ent shades of green indicate the various meadow types and mowing
periods.
On the micro scale, many types of land-use patches display
within-patch production diversity. Arable fields are often used for
mixed cultivation, for example maize, potatoes, and beans. Fields as
well as meadows can host tree stands producing plums and other
fruits used for jam production or distillation of the traditional fruit
brandy, palinka. The spatial diversity between and within patches
interacts with temporal dynamics including field crop rotation, tem-
porary cultivation, and a dynamic meadow-mowing regime.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
6
1
2

216 Lennartsson et al.

Figure 7.7 A grassland landscape, stretching from garden plantations


to fields, meadows, and hilltop pastures, showing a great diversity of
grassland ecosystems. Botiza, Romania.
Photo: Tommy Lennartsson.

Together all these levels of diversity provide favourable condi-


tions for high biological diversity of ‘wild’ flora and fauna, e.g. in
terms of species richness of plants and herbivorous insects. During
summer the landscape is subject to a variety of activities, showing
that biological diversity is formed by a large diversity of land-use
practices. The family field patches and hay meadows are small, of
varying types and spread out in different parts of the village. The
boundaries between them are demarcated by rows of trees, bushes,
and anthills. Each type of meadow and field requires specific prac-
tices for sustainable production of hay and other produce. These prac-
tices are adapted in terms of mowing periods, between-year variation,
presence and frequency of cultivation and fertilisation, bush clearing,
burning of old grass, grazing in spring or autumn, etc. (Dahlström,
Iuga, & Lennartsson 2013).
Garden crops and milk products, especially varieties of tradi-
tional cheese, are traded locally, are partly non-monetary, and are
largely created through social networks.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
7
1
2

Diversity in Ecological and Social Contexts 217

Figure 7.8 The use of the Carpathian agricultural landscape in


Romania is based on mixed pastoralism. The diversity of pastures and
meadows provides feed for grazing livestock, which produce cheese
and other dairy products as well as manure for the arable fields. The
grazers have a key function in the system by transforming grass into
human food. As a by-product, rich biological diversity is formed. Rodna
Mountains, Romania.
Photo: Tommy Lennartsson.

What cannot be observed directly, but instead understood by


interviewing farmers, is that the diversity of material and practical
land use is rooted in a diversity of intangible culture of the landscape
(Figure 7.8). A meadow is clearly more than just a place to cut hay, it
is also an important node of the cultural, economic, social, and even
religious web of the local community (see Chapter 4).

7.6.4.2 Variation and Flexibility


Hay is produced within a variety of land-use types, of which four
are particularly prominent: (1) fields with cultivated fodder,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
8
1
2

218 Lennartsson et al.

(2) fodder-producing leys, (3) annually cut semi-natural meadows,


and (4) irregularly cut semi-natural meadows (Dahlström et al. 2013).
This diversity of fodder-producing areas is partly the result of the
highly variable natural conditions within the village, but also shows
the farmers’ need for flexible resources, and how land use is adapted
to vegetation processes. In combination, the various hay meadows
provide the farmers with a resource that can be adapted and used
differently depending on present and expected needs, the annual sta-
tus of the grassland, the weather, and other factors such as available
workforce or competing duties.
Meadows may change from one type to another as a result of
ecological processes and fluctuating needs of the farmer. An exam-
ple of environmental influence is when annual mowing eventually
depletes the soil nutrients, which leads to reduced production and
sometimes also a vegetation succession towards low-productive and
unpalatable vegetation. Type 3 meadows (see list earlier in this sec-
tion) may require actions such as ploughing or fertilising, which then
turn them into ley meadows for longer or shorter periods. Type 3
meadows may also be improved by periods of rest when no mowing
takes place, and thus turned into type 4 meadows.

7.6.4.3 Coevolution of Practices and Provisioning


Ecosystems – Evolution of Biocultural Diversity
A combination of natural conditions and land use thus shapes the
hay-producing ecosystem in a number of different ways. Ecosystems
and their specific set of land-use practices are a result of repeated
feedback between land use as an ecological process and the response
of the ecosystem. They ‘co-evolve’, to use one ecological term, and
the shaping of the agricultural landscape is a process of ‘niche con-
struction’, to use another such term (see Chapter 6). The linked diver-
sities of land use and ecosystem produce a rich biocultural diversity
which is shaped by the need for and possibility of extraction of a
biological resource, in this case hay. Land-use practices represent the
immediate link between societal needs and ecosystems’ provisioning

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
9
1
2

Diversity in Ecological and Social Contexts 219

potential and can be thus seen as ‘the glue’ between humans and
nature.
This interaction is influenced both by natural and cultural fac-
tors, and together these factors constitute a second, more indirect
interface between humans and the environment. The main natural
factors include weather variations, successional processes in the veg-
etation, and nutrient cycling. From a nutrient perspective, hay mead-
ows are connected to arable land, gardens, and pastures via livestock
and its manure. Some of the cultural factors influencing land use
include:

• Labour availability in relation to need. Labour availability is influenced


by, e.g. family size, social networks, and competing duties. Labour
needs depend on, e.g. the minimum effort needed for harvesting,
maintaining, and restoring different types of provisioning ecosystems, as
well as the ecosystems’ distance to the farm, and expectations of future
resource needs.
• Alternative land use (e.g. grazing, cultivation, or letting land to a
neighbour), related to market, traditions, and social expectations.
• The need for products, which is determined by, e.g. family size and other
sources of income.
• Religious celebrations and traditions related to work-free days.

7.6.4.4 Historical Perspectives


In Western Europe, semi-natural grasslands are generally regarded
as remnants of pre-industrial agricultural landscapes and land-
use regimes. Understanding historical land use is crucial for the
future management of grasslands of high conservation value (e.g.
Oppermann et al. 2012). The Romanian Carpathian landscape is
rich in grassland species, which are threatened with extinction in
many other parts of Europe because of changes in land use (Helldin &
Lennartsson 2007). Above all, it is a low-input agriculture using a
minimum of artificial fertilisers, chemical weeding, and fossil fuels.
The methods and tools employed are largely manual and labour-
intensive, such as scythe-mowing and herding. The land-use practices
employed in the Romanian Carpathians include techniques found in

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
0
2

220 Lennartsson et al.

nineteenth-century historical records from other parts of Europe, and


the low degree of agricultural rationalisation implies that the bio-
cultural heritage from earlier agricultural traditions still exists. In
view of the living biocultural diversity, as well as that from the past,
this landscape can be seen as an historical archive, highly important
for the understanding of agricultural history in many other parts of
Europe (Helldin & Lennartsson 2007).
This ‘archive’ contains a great deal of evidence of earlier forms
of land use, thus indicating changing practices over time. This evi-
dence cannot be interpreted by applying direct relationships between
ecosystem and land-use practices, but a thorough understanding of
political, cultural, and economic factors in a historical context is also
needed.
This can be illustrated by the influence of communist policies,
which is strong both in peoples’ minds and in the landscape. Through
the collectivisation and nationalisation of land during the commu-
nist period of 1945–89, land-use types were rearranged, landmarks
removed, and agriculture often intensified using chemical fertilisers
and pesticides (Verdery 2003). Even villages which were never collec-
tivised, such as Botiza, were strongly influenced by the communist
land-use policies. One such policy was the nationalisation of forests,
which removed many wooded meadows and pastures from the vil-
lage resources and led to intensified use of other types of grassland.
All villages were required to deliver quotas of cultivated crops and
livestock produce. According to villagers, this caused a peak in arable
cultivation in Maramureș in the 1980s. Immediately after the revolu-
tion in December 1989, land began to be returned to former owners
or their relatives. This process has led to considerable changes in land
use (Verdery 2003).
One effect of the more open economy after 1989 has been the
steady decrease in livestock, not least in the mountains. This has led
to reduced demand for hay meadows and pastures. Similarly, as flour
has become more accessible, the cultivation of cereals for human
consumption has decreased in the mountain villages. As a result,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
1
2

Diversity in Ecological and Social Contexts 221

fields can now be used for hay production (as discussed earlier in this
chapter), while meadows situated on village peripheries can be used
for grazing. Faraway meadows and pastures are no longer in use since
land is available at a more convenient distance. These changes have,
on the whole, caused a ‘contraction’ of the agricultural landscape.

7.7 Some Implications and Extensions of


Biocultural Diversity
In the following sections, aspects of natural and cultural diversity
in the context of biocultural diversity are discussed. To begin with,
the examples used in this chapter demonstrate a number of mecha-
nisms by which human society is influenced by biological diversity
and vice versa.

• Land-use systems based on a diverse use of multiple resources develop a)


where the natural resources are scarce due to low productivity and short
growing seasons (Scandinavian reindeer husbandry), b) where the resource
is highly variable, seasonally or between years (reindeer husbandry), and
c) where the landscape is topographically and ecologically heterogeneous
(Romanian mountain farms).
• Deliberate enhancement of biological diversity, in terms of crop
diversity, reduces risk and may be preferred over maximising income and
minimising costs (eighteenth-century Pennsylvania yeomen).
• The use of diverse natural resources needs a diversity of adapted land-
use practices, techniques, and tools that co-evolve with the semi-natural
ecosystems partly shaped by land use (Romanian farming).
• The use of common resources requires a diversity of institutions at
different levels of society to ensure a fair and sustainable, as well as
flexible use of land (Swedish nineteenth-century shielings, reindeer
husbandry).
• In order to obtain a large enough resource in spatially and temporally
variable ecosystems, each household needs diverse land patches; the
property rights being connected to a diversity of institutions and social
networks (Romanian agriculture).
• Diversification of resource use requires an extensive knowledge diversity
(all examples).

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
2

222 Lennartsson et al.

• The local use of ecosystem resources is affected by several non-local


aspects of society, such as collectivisation, EU agricultural policies
(Romanian agriculture), and national policies for forestry, tourism etc.
(reindeer husbandry).

The examples discussed in this chapter refer to systems that contain


several other types of cultural diversity than those explicitly men-
tioned, such as linguistic diversity, market, legislation, spiritual rela-
tionships with nature, etc. Together, all these aspects of biological
and cultural diversity indicate mechanisms for biocultural diversity,
i.e. for the diversity of human-environment interaction.

7.7.1 Correlations versus Mechanisms


Areas with high levels of cultural diversity that geographically over-
lap with high biological diversity are frequently found in tropical and
subtropical regions, e.g. in South and Central America, the Pacific,
South East Asia, and West-Central Africa (Maffi & Woodley 2010).
Such correlation-based biocultural diversity was the first aspect of
biocultural diversity to be described (e.g. Maffi 2001; Moore et al.
2002). Attempts have even been made to quantify this correlation
by combining the cultural and biological diversity of a certain place,
such as a country, using the following indicators: the number of
languages, ethnic groups, religions, bird and mammal species (com-
bined), and plant species in relation to the size of the area and the
human population (Loh & Harmon 2005).
The interpretation and implications of such correlations are not
immediately obvious, and they raise a number of questions. Clearly,
high biodiversity in regions such as the tropics seems to foster a rich
cultural diversity, in various degrees related to the biodiversity, e.g.
in terms of language and other expressions of resource use. But to
what extent does the correlation represent a two-way relationship,
i.e. the effects of cultural diversity on biodiversity? Furthermore,
can we view these regions as hotspots for biocultural diversity (e.g.
Wilcox & Duin 1995), more valuable than temperate regions, which
for biogeographic reasons are less rich in species? In biodiversity

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
3
2

Diversity in Ecological and Social Contexts 223

conservation, as a comparison, such reasoning is generally not


applied, mainly because different types of biological diversity are not
replaceable, i.e. loss of a few species in one species-poor region or
habitat is not compensated by preservation of several other species in
a species-rich region. The last question relates to the scale at which
biological and cultural diversity occur. How much of the entire rain-
forest biodiversity – which may contribute to the high biocultural
diversity index – is used and influenced by the local communities,
and how much would look the same also in the absence of humans?
The question is if our view of biocultural diversity would change
with an increased focus on causal relationships between culture and
environment. In agricultural landscapes, such as that of Romania dis-
cussed in this chapter, most of the biological and cultural diversity
is interrelated. The number of causal relationships between biodiver-
sity and cultural expressions can be expected to be very high even if
the number of species is low compared to rainforest regions.
Subsequent research has increasingly emphasised the impor-
tance of causal relationships between cultural and biological diver-
sity on local levels (Loh & Harmon 2005; Maffi & Woodley 2010).
Knowledge of mechanisms and causal links is necessary in order to
move beyond descriptions and analogies, as observed concerning the
evolution of both biological (e.g. Mayr 1991) and cultural diversity
(Durham 1976). In order to identify such mechanisms, we may need
to focus on those segments of the biological and cultural diversities
with reciprocal influences, and which in many cases are prerequisites
for one another.
The study of mechanisms and significances of biocultural
diversity is necessary if the concept is to play a role in policy-making
and other applied contexts. The application of biocultural diversity
will be addressed in the final section of this chapter.

7.7.2 Biocultural Diversity at Different Levels


The examples examined in this chapter show that different intersects
between culture and environment can be identified, and that these

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
4
2

224 Lennartsson et al.

represent different levels of biocultural diversity. The most direct


intersect is where human practice and the environment meet in
physical terms, mainly through the land-use practices used to man-
age the ecosystems. These practices are in turn influenced by cul-
tural and societal expressions of resource use, such as the diversity of
markets, social interactions, gender relations, traditions, and norms
and values, and the influence represents a more indirect intersect
between humans and the environment. Both intersects are bidirec-
tional, i.e. the local communities change the natural conditions of
the landscape, while at the same time adapting to these conditions.
In Romanian mountain agriculture, the first intersect constitutes the
interactions between the ecological-topographical complexity and
land-use diversity, while the second intersect addresses the relation-
ships between land use and the societal complexity of the densely
populated agrarian villages.
We need to understand the mechanisms that influence both
direct and indirect human-environment links. This is necessary in
order to detect and interpret biocultural diversity, and to understand
social-ecological interactions in general. In the following section, we
will examine how analysis of drivers for biocultural diversity can be
one way of identifying such mechanisms.

7.7.3 Drivers of Biocultural Diversity


By understanding what influences evolution or loss of biocultural
diversity, it is possible to identify important causal links between
society and environment and the values and advantages, or perhaps
disadvantages, of a rich biocultural diversity. This chapter contains
examples of different types of such drivers for the evolution of bio-
cultural diversity.
One is pronounced spatial and temporal biological complex-
ity. In the Romanian Carpathians, spatial biological complexity is
caused by fine-scale geo-topographical variation, the temporal com-
plexity by weather variation and successional ecological processes
in semi-natural ecosystems. Diversified resource use is essential for

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
5
2

Diversity in Ecological and Social Contexts 225

the adaptation of spatial as well as temporal biological complexity.


For example, agricultural pastoralism in seasonal climates requires
a subdivision of the land into summer pastures and hay meadows
for winter fodder. Reindeer husbandry, in contrast, is based on ani-
mals that are adapted to eating lichen in winter and do not require
harvested winter fodder. Biological complexity is nevertheless an
important driver for biocultural diversity since winter and summer
forage are found in ecosystems located far apart. In addition, diver-
sity of summer and winter pastures is needed, since changeable
weather makes each pasture ecosystem subject to annual variation.
Reindeer herders therefore need extensive knowledge of their eco-
systems’ resources, their interplay with climate, weather, and the
semi-domesticated reindeer. The territories, as well as the range of
ecosystems and resources used, are enormous.
In other cases, the ecosystems and their use are less com-
plex, and instead the evolution of biocultural diversity seems to
be mainly driven by complex cultural conditions. The example of
Swedish historical summer farming shows that the villagers’ col-
lective use of pastures at their summer farms required an intricate
diversity of institutions and regulations that largely shaped the bio-
cultural diversity. The Pennsylvania yeomen provide another exam-
ple of how biocultural diversity was driven by societal conditions.
Early American land-use ecology was to some extent opposite to
that of later centuries and also the European equivalent. Labour in
the colonies was very expensive, if available at all, but land was
almost free for the taking. An average landowner held hundreds of
acres of excellent farmland and meadows, but had very little help
to work it. As a consequence, the European system of fertilised per-
manent fields was replaced by shifting cultivation like that used by
slash-and-burn horticulturalists. Many of the traditional measures
of productivity – like corn to horn and per-seed ratios – are ill-suited
for people not looking to maximise production per area. Instead, a
more complex measure, that of production per unit of labour, has
to be used.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
6
2

226 Lennartsson et al.

It is reasonable to assume that local biocultural diversity has


evolved due to local communities’ continuous adaptation to its
(changing) ecological and cultural environment, with sustainability
as an important incentive. The examples in this chapter illustrate
how some mechanisms of biocultural diversity may be linked to sus-
tainability. In historical Pennsylvania, crop diversity needed diver-
sity of labour, or more precisely, diversity of knowledge and labour
skills. Since crop diversity increased the sustainability in unpredicta-
ble conditions, we see that sustainability is linked to both crop diver-
sity and labour/knowledge diversity. This link implied an increased
cost for farmers pursuing high sustainability, both in terms of costs
of skilled and more expensive labour, and in terms of reduced income
from the major market crop, wheat.
In Swedish historical summer farming, the diversification of
institutions and locally adapted regulations increased the resilience
of the system and helped it survive through 300 years of substan-
tial societal change. Finally, however, the system collapsed – appar-
ently having lost its sustainability. It was societal change that finally
made the institutional diversity reach the limit of its capacity. At
least three major conditions influenced this: (1) ecological, where the
cultivated fodder in the villages replaced the importance of resource
diversity in terms of a variety of semi-natural pastures and bog mead-
ows at the summer farms, (2) economic, where new textile products
replaced leather and wool, i.e. a falling demand for the products of
the system, (3) policy-related changes, where the emerging forestry
industry, supported by several national policy measures, competed
for forest resources, which resulted in a gradual loss of forest pasture,
which was the basal ecosystem resource for the summer farms.

7.7.4 Biocultural Diversity and Historical Ecology


Historically, local land-use diversity and ecosystem diversity have
developed simultaneously, and have been closely linked through
dynamic feedback mechanisms. Landscapes and resources have
been shaped by land use together with the knowledge of how to
use the available resources. Even when conditions change, a certain

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
7
2

Diversity in Ecological and Social Contexts 227

inertia means that both environment and culture are still to some
extent characterised by their past. This implies that both cultural
and biological diversity constitute important elements of cultural
heritage and historical source materials (see further Chapter 9), and
that neither culture nor nature can be understood without historical
knowledge.
Two of the examples discussed in this chapter are based on
historical sources, while the other two are based on the present. By
interpreting these in terms of biocultural diversity, they may be used
as tools for the analysis of human-nature relationships in both histor-
ical and present-day land-use systems, especially on the local scale.
This will in turn contribute to the understanding of both biological
systems and societies.
One example of a historical-ecological application of biocul-
tural diversity knowledge is provided by the Pennsylvania yeomen.
Since the emergence of the agricultural reform movement in the
seventeenth century, landlords and agricultural reformers have
lamented the loss of short-term profits by farmers and yeomen and
expressed frustrations with practices seen as backwards or primitive
(Prothero 1961). Knowledge of the relationships between biocultural
diversity and risk mitigation, however, show that these practices
are not foolish, but rather that they maintain or improve biocultural
diversity as a way of managing the social and environmental risks
of farming.
There are four fundamental strategies of risk management
(Winterhalder, Lu, & Tucker 1999: 339). In terms of agricultural pro-
duction, cultivators can manage risk by spreading potential loss of
production over (1) space, by scattering cropped fields, (2) time, by
cadence of planting and storage, (3) products, through diversification,
and (4) producers, by pooling labour and sharing resources. McCloskey
(1975, 1976, 1989, 1991) has championed the view that the inefficien-
cies of scattered plots and communal production in the open-field
system of European manors were a rational response by peasants to
minimise risks to household sustainability. McCloskey’s view rests
on Bloch’s (1966) argument that scattered plots within French open

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
8
2

228 Lennartsson et al.

fields were an insurance against natural and human-made disasters.


This is widely supported in studies of smallholder agriculture by
historians, anthropologists, and economists (see McCloskey 1976,
1989; Bentley 1987; Goland 1993; and Netting 1993 for literature
reviews). McCloskey’s analysis of yields and rents from the thir-
teenth to the nineteenth century reveal that by scattering their fields,
typical English midland farmers traded one tenth of their potential
income for a 50 per cent reduction in risk of catastrophic loss of yield
(McCloskey 1976: 132–65; McCloskey 1989: 5–51).
Another example is the comparative analysis of data derived
from historical and current systems. In the Romanian mountains, it
is possible to observe, interpret, and receive direct information about
land-use diversity and biological diversity in these semi-natural eco-
systems. This contemporary information can be analysed in order
to identify cause-response mechanisms between land use and eco-
systems, and hypotheses can be verified by field studies. We can,
for example, study the response of certain species of meadow plants
to various mowing practices, or the need for maintenance practices
in different meadow types due to vegetation succession. It can be
assumed that certain types of environmental conditions influence
certain species in ways that are independent of time and place (cf.
Scott 1963), and, therefore, that certain ecological processes that
can be observed in contemporary Romania may have been preva-
lent historically, in Romania as in other regions with similar land-
use practices and landscapes (see Chapter 9). By understanding such
relationships, there is an increased potential for interpreting in new
ways historical sources as well as ecological information. Such
understanding is often crucial, not least for biodiversity conservation
(Beaufoy et al. 1994; Plieninger et al. 2006).

7.7.5 Applied Biocultural Diversity


This chapter ends with a discussion about two different types of land-
scape and how the concept of biocultural diversity may contribute
to a better knowledge base for policy and governance development.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
9
2

Diversity in Ecological and Social Contexts 229

The examples referred to in this chapter all deal with local


communities and landscapes where biocultural diversity provides
links between environments and the people who use them and live
there. Such links are clearly visible in agricultural landscapes as
interdependence between rural cultural diversity and semi-natural
biological diversity, mainly linked by land use and its ecological
imprints. Agrarian landscapes are acknowledged as being cultural,
or ‘domesticated’ as discussed in Chapter 6. Yet, as the Romanian
example illustrates, current policies, although aiming at preserving
both biodiversity and cultural heritage, risk causing adverse effects
by inhibiting the use of important traditional knowledge and prac-
tices, thereby destroying ecosystems, limiting sustainable resource
use, and eliminating cultural links to the landscape (Dahlström et al.
2013). In the Romanian case, this is clearly a result of insufficient
awareness of the importance of a number of key components of the
human-nature relationship, mainly those relating to hay meadow
management and land-use configuration. The concept of biocultural
diversity can provide a framework for identifying such key compo-
nents, by indicating how cultural and biological diversity are linked,
and which elements of cultural diversity are necessary for maintain-
ing biodiversity (cf. McNeely 2000).
Other examples in this chapter describe biocultural diversity
in alpine and boreal forest landscapes. These are often viewed as
natural (Ellis 2011), as natural processes have had a large impact on
them. This view, however, omits biocultural diversity and important
human-nature links, which frequently leads to flawed conclusions
regarding land use and landscape history. ‘Pristine nature’ often con-
stitutes the baseline for conservation and policy in forest and alpine
landscapes (see Chapter 5), which may threaten both biodiversity
and local culture, e.g. reindeer herders and transhumant farmers.
In contrast, research within the social sciences and the humanities
regularly claims that alpine and boreal landscapes are ‘cultural land-
scapes’ (Plumwood 2006; Quétier et al. 2010). Neither view, however,
provides an unambiguous baseline for landscape management, unless

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
0
3
2

230 Lennartsson et al.

the importance of different types of human impact on the landscape


and its cultural and biological diversity, as well as the importance of
natural processes, are underscored. Understanding the mechanisms
behind biocultural diversity can provide a structural framework for
describing the interplay between cultural and natural processes in
the landscape.
In conclusion, understanding the enduring interdependence of
local communities and ecosystems – those of the past, current, and
future – requires an integrative framework and new terminology such
as that of biocultural diversity. Such an approach needs a nuanced
understanding of the different components, how they are integrated,
and the importance of their interaction. Such analysis can reveal the
causes of change in biocultural diversity, which in turn can explain
important components of human-environment relationships and are
essential for designing a diverse future.

References
Barthel, S., Crumley, C. L., & Svedin, U. (2013). Bio-cultural refugia: combating
the erosion of diversity in landscapes of food production. Ecology and Society,
18, 71–88.
Beaufoy, G., Baldock, D., & Clark, J. (1994). The Nature of Farming: Low Intensity
Farming Systems in Nine European Countries. London: Institute for European
Environmental Policy.
Bentley, J. W. (1987). Economic and ecological approaches to land fragmentation: in
defense of a much-maligned phenomenon. Annual Review of Anthropology,
16, 31–67.
Biong Deng, L. (2010). Livelihood diversification and civil war: Dinka communi-
ties in Sudan’s civil war. Journal of Eastern African Studies, 4, 381–99.
Bloch, M. (1966). French Feudal History, translated by Janet Sondheimer.
Originally published (1931) as Les caractères originaux de l’histoire rurale
française. Berkley: University of California.
Bloch, M. (2015). The Historian’s Craft. Manchester: Manchester University Press.
Originally published (1949) as Apologie pour l’Histoire ou Métier d’Historien.
Paris: Armand Colin.
Block, S. & Webb, P. (2001). The dynamics of livelihood diversification in post-
famine Ethiopia. Food Policy, 26, 333–50.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
1
3
2

Diversity in Ecological and Social Contexts 231

Blomkvist, P. & Larsson, J. (2013). An analytical framework for common-pool


resource–large technical system (CPR-LTS) constellations. International
Journal of the Commons, 7, 113–39.
Brännlund, I. & Axelsson, P. (2011). Reindeer management during the colonization
of Sami lands: a long-term perspective of vulnerability and adaptation strate-
gies. Global Environmental Change, 21, 1095–1105.
Bromley, D. W. (1991). Environment and Economy, Property Rights and Public
Policy. Oxford: Basil Blackwell Ltd.
Chen, M. A. (1991). Coping with Seasonality and Drought. New Delhi: Sage
Publications.
Cocks, M. (2006). Biocultural diversity: moving beyond the realm of ‘indigenous’
and ‘local’ people. Human Ecology, 34(2), 185–200.
Couture, J. (2000). Native studies and the Academy. In G. J. Sefa Dei, B. Hall, & D.
Goldin Rosenberg, eds., Indigenous Knowledges in Global Contexts: Multiple
Readings of Our World. Toronto: University of Toronto Press, pp. 157–67.
Dahlström, A., Iuga, A., & Lennartsson, T. (2013). Managing biodiversity rich
hay-meadows in the EU: a comparison of Swedish and Romanian grasslands.
Journal of Environmental Conservation, 40(2), 194–205.
Danell, Ö. (2000). Status, directions and priorities of reindeer husbandry research
in Sweden. Polar Research, 19, 111–15.
Dasmann, R. F. (1991). The importance of cultural and biological diversity. In M.
L. Oldfield & J. B. Alcorn, eds., Biodiversity, Culture, Conservation, and Eco-
development. Boulder, CO: Westview, pp. 7–15.
Davies, E. (1941). The patterns of transhumance in Europe. Geography, Journal of
the Geographical Association, 117, 155–68.
Descola, P. & Pálsson, G. (1996). Introductory chapter. In P. Descola & G.
Pálsson, eds., Nature and Society. Anthropological Perspectives. London,
New York: Routledge, pp. 1–21.
Dietz, T., E. Ostrom, & P. C. Stern. (2003). The struggle to govern the commons.
Science, 302, 1907–12.
Doak, D. F., Bigger, D., Harding, E. K., Marvier, M. A., O’Malley, R. E., & Thomson,
D. (1998). The statistical inevitability of stability-diversity relationships in
community ecology. The American Naturalist, 151, 264–76.
Durham, W. H. (1976). The adaptive significance of cultural behaviour. Human
Ecology, 4, 89–121.
Dynesius, M. & Jansson, R. (2000). Evolutionary consequences of changes in spe-
cies’ geographical distributions driven by Milankovitch climate oscillations.
Proceedings of the National Academy of Sciences of the United States of
America, 97, 9115–20.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
2
3

232 Lennartsson et al.

Ellis, E. C. (2011). Anthropogenic transformation of the terrestrial biosphere.


Philosophical Transactions of the Royal Society A, 369, 1010–35.
Ellis, F. (1998). Household strategies and rural livelihood diversification. The
Journal of Development Studies, 35, 1–38.
Ellis, F. (2000). Rural Livelihoods and Diversity in Developing Countries. Oxford:
Oxford University Press.
Ellis, F. & Allison, E. (2004). Livelihood Diversification and Natural Resource
Access. FAO Livelihood support programme Working paper 9.
Elton, C. S. (1958). The Ecology of Invasions by Animals and Plants. London:
Methuens.
Eriksson, O. (2013). Species pools in cultural landscapes – niche construction, eco-
logical opportunity and niche shifts. Ecography, 36, 403–13.
Feeny, D., Berkes, F., McCay, B. J., & Acheson J. M. (1990). The tragedy of the com-
mons: twenty-two years later. Human Ecology, 18(1), 1–19.
Fischer, A. G. (1960). Latitudinal variation in organic diversity. Evolution,
14, 64–81.
Glacken, C. J. (1967). Traces on the Rhodian Shore. Nature and culture in Western
thought from ancient times to the end of the eighteenth century. Berkeley,
Los Angeles, London: University of California Press.
Goland, C. (1993). Field scattering as agricultural risk management: a case
study from Cuyo Cuyo, Department of Puno, Peru. Mountain Research and
Development, 13(4), 317–38.
Gonzalez, A. & Loreau, M. (2009). The causes and consequences of compensatory
dynamics in ecological communities. Annual Review of Ecology, Evolution,
and Systematics, 40, 393–414.
Hardin, G. (1968). The tragedy of the commons. Science, 162, 1243–8.
Harper, J. L. (1969). The role of predation in vegetational diversity. Brookhaven
Symposia in Biology, 22, 48–61.
Hawkins, B. A., Field, R., Cornell, H. W., Currie, D. J., Guégan, J.-F., Kaufman, D.
M., Kerr, J. T., Mittelbach, G. G., Oberdorff, T., O’Brien, E. M., Porter, E. E., &
Turner, J. R. G. (2003). Energy, water, and broad-scale geographic patterns of
species richness. Ecology, 84, 3105–17.
Hector, A., Hautier, Y., Saner, P., Wacker, L., Bagchi, R., Joshi, J., Scherer-Lorenzen,
M., Spehn, E. M., Bazeley-White, E., Weilenmann, M., Caldeira, M. C.,
Dimitrakopoulos, P. G., Finn, J. A., Huss-Danell, K., Jumpponen, A., Mulder,
C. P. H., Palmborg, C., Pereira, J. S., Siamantziouras, A. S. D., Terry, A. C.,
Troumbis, A. Y., Schmid, B., & Loreau, M. (2010). General stabilizing effects of
plant diversity on grassland productivity through population asynchrony and
overyielding. Ecology, 91, 2213–20.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
3
2

Diversity in Ecological and Social Contexts 233

Heggberget, T. M., Gaare, E., & Ball, J. P. (2002). Reindeer (Rangifer tarandus) and
climate change: importance of winter forage. Rangifer, 22, 13–32.
Helldin, J.-O. & Lennartsson, T. (2007). Agricultural landscapes in East Europe
as reference areas for Swedish land management. In V. Surd & V. Zotic, eds.,
Rural Space and Local Development. Cluj-Napoca: Cluj University Press,
pp. 367–70.
Helle, T., Aspi, J., & Kilpelä, S. S. (1990). The effects of stand characteristics
on reindeer lichens and range use by semi-domesticated reindeer. Rangifer
Special Issue, 3, 107–14.
Hodder, I. (2012). Entangled: An Archaeology of the Relationships between
Humans and Things. Oxford: Blackwell Wiley.
Holling, C. S. (1973). Resilience and stability of ecological systems. Annual
Review of Ecology and Systematics, 4, 1–23.
Horstkotte, T. & Roturier, S. (2013). Does forest stand structure impact the dynam-
ics of snow on winter grazing grounds of reindeer (Rangifer t. tarandus)? Forest
Ecology and Management, 291, 162–71.
Horstkotte, T., Sandström, C., & Moen, J. (2014). Exploring the multiple use of
boreal landscapes in northern Sweden: the importance of social-ecological
diversity for mobility and flexibility. Human Ecology, 42, 671–82.
von Humboldt, A. (1808). Views of Nature, Translated by E. Otté. & H. Bohn.
(1850). London: Bohn.
von Humboldt, A. & Bonpland, A. (1807). Essay on the Geography of Plants.
Edited by S. T. Jackson. (2009). Chicago, London: University of Chicago Press.
Ikonen, I. & Lammi, A., eds. (2000). Traditional Rural Biotopes in the Nordic
Countries, the Baltic States and the Republic of Karelia: An International
Seminar and Workshop in Turku May 2–May 4, 2000. Copenhagen: Nordic
Council of Ministers.
IUCN. (2008). Report of the Director General on the Work of the Union since
the IUCN World Conservation Congress, Bangkok, 2004. IUCN World
Conservation Congress, Barcelona, Spain 2008. http://cmsdata.iucn.org/
downloads/cgr_2008_8_dg_report.pdf.
Jochim, M. A. (1981). Strategies for Survival. Cultural Behaviour in an Ecological
Context. New York: Academic Press.
Johnson, J. T. & Murton, B. (2007). Re/placing native science: indigenous voices
in contemporary constructions of nature. Geographical Research, 45, 121–9.
Jörgensen, D. (2015). Rethinking rewilding. Geoforum, 65, 482–8.
Kivinen, S., Moen, J., Berg, A., & Eriksson, Å. (2010). Effects of modern forest
management on winter grazing resources of reindeer in Sweden. Ambio, 39,
269–78.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
4
3
2

234 Lennartsson et al.

Kricher, J. (2009). The Balance of Nature: Ecology’s Enduring Myth. Princeton,


NJ: Princeton University Press.
Larsson, J. (2009). Fäbodväsendet 1550–1920: Ett centralt element i Nordsveriges
jordbrukssystem [Summer Farms in Sweden 1550–1920: An Important
Element in North Sweden’s Agricultural System]. Uppsala: Swedish University
of Agricultural Sciences.
Larsson, J. (2012). The expansion and decline of a transhumance system in Sweden,
1550–1920. Historia Agraria, 56, 11–39.
Larsson, J. (2014a). Boundaries and property rights: the transformation of a
common-pool resource. Agricultural History Review, 62, 20–40.
Larsson, J. (2014b). Labor division in an upland economy: workforce in a
seventeenth-century transhumance system. The History of the Family, 19,
393–410.
Leadley, P. W., Krug, C. B., Alkemade, R., Pereira, H. M., Sumaila, U. R., Walpole, M.,
Marques, A., Newbold, T., Teh, L. S. L, van Kolck, J., Bellard, C., Januchowski-
Hartley, S. R., & Mumby, P. J. (2014). Progress towards the Aichi Biodiversity
Targets: An Assessment of Biodiversity Trends, Policy Scenarios and Key
Actions. Global biodiversity outlook 4, Technical Report, CBD Technical Series
78. Montreal, Canada: Secretariat of the Convention on Biological Diversity.
Lewontin, R. C. (1969). The meaning of stability. Brookhaven Symposia in
Biology, 22, 13–24.
Loh, J. & Harmon, D. (2005). A global index of biocultural diversity. Ecological
Indicators, 5, 231–41.
Loreau, M., Naeem, S., Inchausti, P., Bengtsson, J., Grime, J. P., Hector, A., Hooper,
D. U., Huston, M. A., Raffaelli, D., Schmid, B., Tilman, D., & Wardle, D. A.
(2001). Biodiversity and ecosystem functioning: current knowledge and future
challenges. Science, 294, 804–8.
Lundmark, H., Josefsson, T., & Östlund, L. (2013). The history of clear-cutting
in northern Sweden. Driving forces and myths in boreal silviculture. Forest
Ecology and Management, 307, 112–22.
Maffi, L. ed. (2001). On Biocultural Diversity: Linking Language, Knowledge, and
the Environment. Washington, DC: Smithsonian Institution Press.
Maffi, L. & Woodley, E. (2010). Biocultural Diversity Conservation, a Global
Sourcebook. New York: Earthscan.
Maurer, K., Weyand, A., Fischer, M., & Stöcklin, J. (2006). Old cultural traditions,
in addition to land use and topography, are shaping plant diversity of grass-
lands in the Alps. Biological Conservation, 130, 438–46.
Mayr, E. (1991). One Long Argument – Charles Darwin and the Genesis of Modern
Evolutionary Thought. Harvard University Press.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
5
3
2

Diversity in Ecological and Social Contexts 235

MacArthur, R. H. (1955). Fluctuations of animal populations, and a measure of


community stability. Ecology, 36, 533–6.
MacArthur, R. H. & Wilson, E. O. (1967). The Theory of Island Biogeography.
Princeton, NJ: Princeton University Press.
McCloskey, D. N. (1975). The economics of enclosure: a market analysis.
In N. William Parker & E. L. Jones, eds., European Peasants and Their
Markets: Essays in Agrarian Economic History, Princeton, NJ: Princeton
University Press, pp. 123–60.
McCloskey, D. N. (1976). English open fields as behavior towards risk. Research
in Economic History, 1, 124–70.
McCloskey, D. N. (1989). The open fields of England: rent, risk, and the rate of
interest, 1300–1815. In D. W. Galenson, ed., Markets in History: Economic
Studies of the Past. Cambridge: Cambridge University Press, pp. 5–51.
McCloskey, D. N. (1991). The prudent peasant: new findings on open fields.
Journal of Economic History, 51, 343–55.
McKean, M. (1986). Management of traditional common lands (Iriaichi) in Japan. In
Proceedings of the Conference on Common Property Resource Management.
National Research Council, Washington, DC: National Academy of Sciences,
pp. 533–89.
McNaughton, S. J. (1977). Diversity and the stability of ecological communities: a
comment on the role of empiricism in ecology. The American Naturalist, 11,
515–25.
McNeely, J. A. (2000). Cultural factors in conserving biodiversity. In A. Wilkes,
H. Tillman, M. Salas, T. Grinter, & Y. Shaoting, eds., Links between Cultures
and Biodiversity. Proceedings of the Cultures and Biodiversity Congress.
Yunnan Science and Technology Press, pp. 128–82.
McNeely, J. A. & Keeton, W. S. (1995). The interaction between biological and
cultural diversity. In B. von Droste, H. Plachter, & M. Rössler, eds., Cultural
Landscapes of Universal Value: Components of a Global Strategy. Jena and
New York: Fischer and UNESCO, pp. 25–36.
Menge, B. A. & Sutherland, J. P. (1976). Species diversity gradients: synthesis of the
roles of predation, competition, and temporal heterogeneity. The American
Naturalist, 110, 351–69.
Moen, J. (2008). Climate change: effects on the ecological basis for reindeer hus-
bandry in Sweden. Ambio, 37, 304–11.
Moen, J., Andersen, R., & Illius, A. (2006). Living in a seasonal environment. In K.
Danell, R. Bergström, P. Duncan, & J. Pastor, eds., Large Herbivore Ecology,
Ecosystem Dynamics and Conservation. Cambridge: Cambridge University
Press, pp. 50–70.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
6
3
2

236 Lennartsson et al.

Moen, J. & Danell, Ö. (2003). Reindeer in the Swedish mountains: an assessment


of grazing impacts. Ambio, 32, 397–402.
Moen, J. & Keskitalo, E. C. H. (2010). Interlocking panarchies in multi-use boreal
forests in Sweden. Ecology and Society, 15(3), 17.
Monbiot, G. (2013). Feral: Searching for Enchantment on the Frontiers of
Rewilding. London: Allen Lane.
de Moor, T. (2015). The Dilemma of the Commoners: Understanding the Use
of Common Pool Resources in Long-Term Perspective. New York: Cambridge
University Press.
Moore, J. L., Manne, L., Brooks, T., Burgess, N. D., Davies, R., Rahbek, C.,
Williams, P., & Balmford, A. (2002). The distribution of cultural and biolog-
ical diversity in Africa. Proceedings of the Royal Society of London, 269,
1645–53.
Netting, R. M. C. (1981). Balancing on an Alp. Cambridge: Cambridge University
Press.
Netting, R. M. C. (1993). Smallholders, Householders: Farm Families and the
Ecology of Intensive, Sustainable Agriculture. Stanford, CA: Stanford
University Press.
North, D. C. (1990). Institutions, Institutional Change and Economic Performance.
Cambridge: Cambridge University Press.
North, D. C. (1991). Institutions. The Journal of Economic Perspectives, 5(1),
97–112.
Oppermann, R., Beaufoy, G., & Jones, G., eds. (2012). High Nature Value Farming
in Europe. Ubstadt-Weiher: Verlag Regionalkultur.
Orr, A., Blessings, M., & Saiti-Chitsonga, D. (2009). Exploring seasonal poverty
traps: the ‘six-week window’ in southern Malawi. Journal of Development
Studies, 45, 227–55.
Östlund, L., Zachrisson, O., & Axelsson, A.-L. (1997). The history and trans-
formation of a Scandinavian boreal forest landscape since the 19th century.
Canadian Journal of Forest Research, 27, 1198–1206.
Ostrom, E. (1990). Governing the Commons: The Evolution of Institutions for
Collective Action. Cambridge: Cambridge University Press.
Ostrom, E. (1992). Crafting Institutions for Self-Governing Irrigation Systems.
San Francisco, CA: Institute for Contemporary Studies Press.
Ostrom, E. (2005). Understanding Institutional Diversity. Princeton, NJ: Princeton
University Press.
Ostrom, E., Gardner, R., & Walker, J. (1994). Rules, Games, and Common-Pool
Resources, Ann Arbor: University of Michigan Press.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
7
3
2

Diversity in Ecological and Social Contexts 237

Ouma, C., Obando, J., & Koech, M. (2012). Post drought recovery strategies
among the Turkana pastoralists in northern Kenya. Scholarly Journals of
Biotechnology, 1, 90–100.
Paine, R. T. (1966). Food web complexity and species diversity. The American
Naturalist, 100, 65–75.
Paine, R. T. (1969). A note on trophic complexity and community stability. The
American Naturalist, 103, 91–3.
Panjek, A. & Larsson, J., eds. (2017). Integrated Peasant Economy in a Comparative
Perspective. Alps, Scandinavia, and Beyond. Koper: University of Primorska
Press.
Plieninger, T., Höchtl, F., & Spek, T. (2006). Traditional land-use and conservation
in European rural landscapes. Environmental Science and Policy, 9, 317–21.
Plumwood, V. (2006). The concepts of a cultural landscape, nature, culture and
agency in the land. Ethics and the Environment, 11, 115–50.
Posey, D. A., ed. (1999). Cultural and Spiritual Values of Biodiversity.
Nairobi: UNEP and Intermediate Technology Publications.
Pretty, J., Adams, B., Berkes, F., Ferreira de Athayde, S., Dudley, N., Hunn, E.,
Maffi, L., Milton, K., Rapport, D., Robbins, P., Sterling, E., Stolton, S., Tsing,
A., Vintinnerk, E., & Pilgrim, S. (2009). The intersections of biological diver-
sity and cultural diversity: towards integration. Conservation and Society, 7,
100–12.
Prothero, R. (1961). English Farming: Past Present. 6th ed. Chicago: Quadrangle
Books. First published in 1912 by Longmans, Green & Co. Ltd,
Quétier, F., Rivoal, F., Marty, P., de Chazal, J., Thuiller, W., & Lavorel, S. (2010).
Social representations of an alpine grassland landscape and socio-political
discourses on rural development. Regional Environmental Change, 10,
119–30.
Reardon, T., Taylor, J. E., Stamoulis, K., Lanjouw, P., & Balisacan, A. (2000).
Effects of nonfarm employment on rural income inequality in developing
countries: an investment perspective. Journal of Agricultural Economics, 51,
266–88.
Root, R. B. (1967). The niche exploitation pattern of the blue-grey gnatcatcher.
Ecological Monographs, 37, 318–50.
Roturier, S. & Roué, M. (2009). Of forest, snow and lichen: Sami reindeer herd-
ers’ knowledge of winter pastures in northern Sweden. Forest Ecology and
Management, 258, 1960–7.
Sanders, H. L. (1968). Marine benthic diversity: a comparative study. The American
Naturalist, 102, 243–82.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
8
3
2

238 Lennartsson et al.

Sandom, C., Donlan, J., Svenning, J.-C., & Hansen, D. (2013). Rewilding. In D. W.
Macdonald & K. J. Willis, eds., Key Topics in Conservation Biology 2. Wiley,
pp. 430–51.
Sandström, C., Moen, J., Widmark, C., & Danell, Ö. (2006). Progressing toward co-
management through collaborative learning: forestry and reindeer husbandry
in dialogue. International Journal of Biodiversity Science and Management,
2, 326–33.
Scholl, M. D. (2008). The American Yeoman: An Historical Ecology of Production
in Colonial Pennsylvania. Chapel Hill: University of North Carolina.
Scholl, M. D., Murray, D. S., & Crumley, C. L. (2010). Comparing trajectories
of climate, class and production: an historical ecology of American yeo-
men. In I. Vaccaro, E. A. Smith, & S. Aswani, eds., Environmental Social
Sciences: Methods and Research Design. Cambridge: Cambridge University
Press, pp. 322–48.
Scott, G. H. (1963). Uniformitarianism, the uniformity of nature, and paleoecol-
ogy. New Zealand Journal of Geology and Geophysics, 6, 510–27.
Simpson, G. G. (1964). Species density of North American recent mammals.
Systematic Zoology, 13, 57–73.
Siy, R. Y. (1982). Community Resource Management, Lessons from the Zinjera.
Quezon City: University of the Philippines Press.
Skutnabb-Kangas, T., Maffi, L., & Harmon, D. (2003). Sharing a World of
Difference: The Earth’s Linguistic, Cultural, and Biological Diversity.
Paris: UNESCO, Terralingua, WWF.
Smith, D. R, Gordon, A., Meadows, K., & Zwick, K. (2001). Livelihood diversifica-
tion in Uganda: patterns and determinants of change across two rural districts.
Food Policy, 26, 421–35.
Smith, E. A. (2001). On the coevolution of cultural, linguistic, and biological diver-
sity. In L. Maffi, ed., On Biocultural Diversity. Washington, DC: Smithsonian
Institution Press, pp. 95–117.
Snow, C. P. (1959). The Two Cultures and the Scientific Revolution. The Rede
Lecture. London: Cambridge University Press.
Soulé, M. E. & Noss, R. F. (1998). Rewilding and biodiversity: complementary
goals for continental conservation. Wild Earth, 8, 19–28.
Stern, P. C. (2011). Design principles for global commons: natural resources and
emerging technologies. International Journal of the Commons, 5, 213–32.
Swinnen, J. F. M. (1999). The political economy of land reform choices in Central
and Eastern Europe. Economics of Transition, 7, 637–64.
Tylor, E. B. (1871). Primitive Culture: Researches into the Development of
Mythology, Philosophy, Religion, Language, Art, and Custom Vols. 1–2.
London: John Murray.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
9
3
2

Diversity in Ecological and Social Contexts 239

UNEP. (2007). Global Environment Outlook 4. Nairobi: UNEP.


UNESCO. (2001). Resolutions, vol. 1. Records of the General Conference, 31st
session, Oct.–Nov.
Verdery, K. (2003). The Vanishing Hectare: Property and Value in Postsocialist
Transylvania. Ithaca, NY: Cornell University Press.
Wallace, A. R. (1878). Tropical Nature and Other Essays. New York: Macmillan.
Whittaker, R. H. (1960). Vegetation on the Siskiyou Mountains, Oregon and
California. Ecological Monographs, 30, 279–338.
Whittaker, R. H. (1972). Evolution and measurement of species diversity. Taxon,
21, 213–51.
Wilcox, B. A. & Duin, K. N. (1995). Indigenous cultural and biological diver-
sity: overlapping values of Latin American ecoregions. Cultural Survival
Quarterly, 18, 49–53.
Williams, G. (1977). Differential risk strategies as cultural style among farmers in
the Lower Chobut Valley, Patagonia. American Ethnologist, 4, 65–83.
Winterhalder, B., Lu, F., & Tucker, B. (1999). Risk-sensitive adaptive tactics: mod-
els and evidence from subsistence studies in biology and anthropology. Journal
of Anthropological Research, 7(4), 301–48.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:41, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.007
0
4
2

8 How to Operationalise
Collaborative Research
Elizabeth A. Jones, Anna Westin, Scott Madry,
Seth Murray, Jon Moen, and Amanda Tickner

‘The secret is to gang up on the problem, rather than each other.’

– Thomas Stallkamp

8.1 Introduction: The Necessity and


Benefits of Collaboration in Interdisciplinary
Research
A key component of historical ecology is the dialectical interaction
between humans and their environment (Crumley & Marquardt
1987; Crumley 1994; Balée 1998). The investigation of this relation-
ship includes phenomena such as culture, society, climate, animal
and plant populations, forested and agricultural landscapes, and the
built environment (e.g. cities, roads, canals). Thus, historical ecology,
by definition, requires both an interdisciplinary theoretical frame-
work, and in most cases a collaborative research approach. Such
interdisciplinary work produces many synergistic benefits. The pri-
mary benefit is that collaboration between scholars working within
and across different disciplines generates questions and proposes
answers that would be difficult, if not impossible to produce within
traditional individual disciplinary boundaries (Moen 2015). The use
and integration of different data sets and analytical methods provides
a more complex understanding of phenomena, allowing examination
of a greater number of factors, revealing relationships that would
go undetected when using only one data set and a single method
of analysis or interpretation. Even when examining the same data,
researchers trained in different disciplines interpret the data from dif-
ferent perspectives. When these alternative analytical viewpoints are
shared and integrated, they create a more multifaceted and intricate
240

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
1
4
2

How to Operationalise Collaborative Research 241

picture of the evidence. An additional benefit of interdisciplinary


research is that different, independent data sets can be used to cali-
brate each other or to fill in gaps with analogous information. This
is particularly important when dealing with phenomena that span
longer time frames, or when focussing on the distant past, for which
data sets are typically sporadic and incomplete. Another powerful
advantage of using multiple types of data and comparing independent
analyses is derived from the contradictions that may appear between
different data sets that are frequently based on differing temporal and
spatial scales. The resolution of such contradictions through inte-
grated research often serves to identify and correct errors that would
otherwise go undetected. This process of resolving contradictions
necessitates calibrating, aligning, and linking data sets, and through
this process, new questions and insights are often generated, which
have the potential to significantly advance the research and at times
move it in completely new directions.
Although it is widely accepted that the complex phenomena
studied in historical ecology require interdisciplinary approaches for
effective explanation, the conceptual and practical aspects of inter-
disciplinary collaboration and integration can be daunting, and a
majority of researchers are not trained to work in this way. The study
of culture and society has traditionally been the purview of the social
sciences (sociology, anthropology, linguistics, psychology, etc.) and
humanities (history, culture studies, literature, etc.), while the study
of environment is more firmly rooted in the natural sciences (biol-
ogy, chemistry, geology, meteorology, etc.). Some disciplines, such
as ecology, geography, agricultural history, and archaeology have,
due to the nature of their data, bridged the boundaries between the
natural sciences, the social sciences, and the humanities. Even so,
specialists within these disciplines often adhere to narrower, tradi-
tional disciplinary perspectives. Substantial theoretical challenges
exist in addressing the different assumptions that the humanities,
social sciences, and natural sciences hold about the world and each
other, including varying ideas about the very nature of the process

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
2
4

242 Jones et al.

of scholarly inquiry. Many scholars have discussed the theoretical


advantages and disadvantages of interdisciplinary research (Crumley
2005, 2007a; Chettiparamb 2007; Strang 2009; Boix-Mansilla 2010;
Frodeman, Klein, & Mitcham 2012). This chapter is not intended to
be a summary of that body of work. Other chapters in this volume
discuss theoretical approaches in historical ecology, including the
use of common concepts (Chapter 6) and the links between human
societies and biological diversity (Chapter 7). This chapter instead
addresses the social aspects of collaboration, in order to provide some
practical considerations and suggestions for collaborative interdisci-
plinary work, based on examples from the authors’ experience in car-
rying out historical ecology research.

8.2 Common Research Questions


From the outset, scholars embarking on collaborative research in his-
torical ecology, with a focus on a particular landscape or environ-
mental issue, must develop key questions to guide their research.
Research initiated by questions developed within one discipline,
but which requires input from other disciplines, is called multi-
disciplinary research. Such research is conducted independently
within the traditional disciplinary boundaries, but results are sum-
marised and synthesised after the independent analyses are com-
plete. Interdisciplinary research is more integrative, but note that
no universal agreement exists on what such an approach entails. Our
vision is that questions and the overall research design should be
developed collaboratively, with the questions being equally relevant
to all research partners. By this we mean that collaborators must be
able to address common questions through their own particular data,
using the methods and standards of their specific discipline, and the
answers to the research questions need to be as meaningful within
the individual disciplines as they are for the overarching, integrated,
interdisciplinary interpretation. Finally, transdisciplinary research
transcends the disciplinary boundaries and thus creates a holis-
tic approach that is more than just the sum of the parts. In other

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
3
4
2

How to Operationalise Collaborative Research 243

words, it aims to create more inclusive theoretical perspectives and


analytical methods that bridge disciplinary boundaries. In striving
for more inclusive perspectives, the transdisciplinary approach also
aims to break down the divide between academic and non-academic
knowledge (see the discussion about TEK later in this chapter). ‘The
common words for multidisciplinary, interdisciplinary and transdis-
ciplinary are additive, interactive, and holistic, respectively’ (Choi &
Pak 2006: 351). In this chapter, we wish to promote collaborations
developed in the intersections between disciplines, where questions
and methods are developed through equal ‘ownership’ and mutual
interest. These will lead to a deeper engagement of the collabora-
tors and to a greater understanding of one another’s perspectives
and expertise, and will also produce results that address a wider
range of factors and how they interact in the production of complex
phenomena.
Formulating such overarching research questions is a time-
consuming and complex process. The available data sets must be
based on comparable and appropriate temporal and spatial scales in
order to support the research questions and enable effective integra-
tion of diverse data. The development of common and meaningful
questions requires time and familiarity with one another’s data. This
is achieved through discussions and formal presentation of the indi-
vidual research activities. The guiding research questions need to be
open to modification during the course of the research as new results
or relationships are revealed, or as new data become available.

8.3 Stakeholder Involvement


Research in a particular landscape setting should also include some
level of collaboration or consultation both with local inhabitants –
those who make a living from the land as well as those who simply
reside there – and non-local stakeholders with ties to the location,
such as local/regional government representatives or even groups with
national or global interests tied to local areas of national economic
or cultural significance. The local inhabitants may be involved in

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
4
2

244 Jones et al.

Figure 8.1 Farmers, researchers, and government officials meet in the


field to discuss the history and future management of the landscape.
Hjälmö farm, county of Stockholm, Sweden.
Photo: Anna Westin.

various ways, e.g. as collaborators in the development of the research


questions, or as providers of different types of data (see Figures 8.1
and 8.2). Such aspects are particularly relevant for ‘knowledge mobi-
lisation’ relating to environmental resources involving local com-
munities, scientists, and governmental entities (Whyte 2013). At the
very least, people must be informed about ongoing research in their
locality, particularly when conducted with and among people who
have been historically disenfranchised in political, economic, and
social terms, or those who have been excluded from providing input
on other types and scopes of research conducted in and around their
communities.
Local participation is a key aspect of how ethnographers gener-
ally approach fieldwork, and their discipline offers a well-established
methodological model for engaging local participation in academic
research in a transparent and ethical manner (Jaarsma 2002; Madison
2012; Okely 2012; O’Reilly 2012; Schensul & Lecompte 2013). When

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
5
4
2

How to Operationalise Collaborative Research 245

Figure 8.2 Researchers studying a farm archive in the house where the
documents were written, together with descendants of the document
author. Hyttbäcken farm, county of Dalarna, Sweden.
Photo: Anna Westin.

these local and non-local stakeholders are brought into the collabo-
ration, research questions need to relate to them at some level, e.g.
pertaining to their cultural history, identity, or current land-use or
environmental issues that are important to them. Local involvement
can help refine, reorient, or retool research questions in consequen-
tial and beneficial ways. It can also provide a deeper understanding
of the complexity of relationships that communities have with their
landscape. Such interaction may introduce influential aspects of the
socio-ecological system not initially apparent to researchers, such
as traditions and collective wisdom influencing land use, traditional
ecological knowledge (TEK), and religious practices (including both
local place-based rituals and broader religious world views). The tra-
ditional knowledge and inherited cultural practices both reflect and
shape a community’s relationship with the land (Chapter 4).
As discussed previously, transdisciplinary research breaks
down not only boundaries among academic disciplines, but also

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
6
4
2

246 Jones et al.

boundaries between professional researchers and non-professionals


who are stakeholders in the research. Working with local people as
co-researchers can, just like interdisciplinary research, generate new
questions and results otherwise unobtainable (Figure 8.2). Alongside
these advantages, working with the local population presents chal-
lenges similar to those faced when working with professional col-
leagues discussed in detail later in this chapter. Many aspects and
procedures of research that professionals take for granted will not be
immediately comprehensible to local participants, making it impor-
tant to initially clarify the purpose of the research, roles and responsi-
bilities, funding, publication, rights to data, and details of the working
process. Professionals must be aware of the often unspoken power
relationships between academics and laypeople, which may inhibit
open communication, and they must endeavour to rid the project
of prejudices regarding what is considered ‘real’ or valid knowledge,
i.e. academic knowledge versus local knowledge (or traditional eco-
logical knowledge, TEK). Ideally, a local stakeholder group should
be consulted at the beginning of the project on research questions,
methods, etc. Local involvement can often be initiated through coop-
eration with local entities such as environmental advocacy groups,
social clubs, or retirement communities. Results should be shared
with stakeholders in a manner that is relevant to their interests and
understandable to them, rather than by simply providing copies of
academic papers, which may be irrelevant or difficult to comprehend.
This means that research results may often need to be presented in
various formats adapted for the different audiences. Depending on
the nature of the research, reporting back to the local population may
also require formal presentations, e.g. at a municipal council meet-
ing. Reporting to stakeholders provides the researchers with oppor-
tunities for accountability to the community in which the research
was conducted.
One such example, derived from the work of the authors, is
HagmarksMISTRA, a Swedish transdisciplinary research programme
entitled ‘Semi-natural grasslands – ecology and economy’, conducted

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
7
4
2

How to Operationalise Collaborative Research 247

Figure 8.3 The logotype for the HagmarksMISTRA research


programme, reflecting different disciplines and perspectives used.
Swedish Biodiversity Centre.

2001–8 (Figure 8.3). HagmarksMISTRA focussed on the biodiver-


sity of rich semi-natural grasslands, through combined research on
ecology, land-use history, economy, agrarian production, and social
aspects. The project research relied on frequent communication with
farmers and other stakeholders in order to create questions and sug-
gest management regimes that were not only ecologically ‘correct’,
but also practical for the farmers (Figure 8.1). In addition to scientific
publications, such as PhD theses and scientific reports, the project
produced annual reports aimed at a wider audience, as well as popular
articles on important results. The results in the annual reports were
either previously published in scientific papers or presented in such
a way that they did not compromise future scientific publication

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
8
4
2

248 Jones et al.

(i.e. meeting the requirement that the content had not been previ-
ously published). The final publication was a popular book that pre-
sented the combined research of the entire project. An experienced
journalist, well acquainted with the field of research, reviewed the
published work and interviewed all the researchers (Olsson 2008).
The results were presented in language easy to comprehend for the
lay reader with rich illustration and were consequently appreciated
by a wide audience. Each chapter stated which scientific research
it was based on. The end result satisfied all parties. The collabora-
tors also arranged and participated in scientific and stakeholder
conferences, radio interviews, and open lectures at the study sites
for the local population. The research results had concrete impacts
on policy-making after presentations to and discussions with the
Swedish Board of Agriculture, as changes in regulations of the agri-
environmental programme concerning semi-natural grasslands.

8.4 Building Trust: Communication,


Leadership, and Followership
It is essential for members of a collaborative research project to main-
tain regular face-to-face contact, not only to share data and individual
progress, but also to allow for reflection and exploration of ideas in
ways that are impossible through e-mail exchanges and other indi-
rect forms of communication. Frequent meetings also promote the
candid communication and trust building that is essential for a free
exchange of data and results. Informal activities, such as dinners
and drinks, further serve to break down communication barriers.
The casual conversation that takes place during these social times
increases familiarity with co-researchers, and also often stimulates
the creative flow of ideas beneficial to the research because conversa-
tion can roam free beyond the confines of a meeting agenda.
In some disciplines, such as archaeology and geology, research-
ers are more accustomed to collaborating with a variety of specialists
having differing disciplinary perspectives. In other disciplines, such
as history, literature, and cultural anthropology, solitary researchers

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
9
4
2

How to Operationalise Collaborative Research 249

are more common, and they may not be as comfortable with collabo-
rative frameworks. To enable a free exchange of ideas, collaborative
groups need to be egalitarian, that is, members should feel comfort-
able in expressing their ideas freely regardless of status and position.
It is preferable to establish rules at the outset concerning the expec-
tations for interaction and communication, data sharing, and pub-
lishing. Guidelines for group membership should also be established,
e.g. the addition of new members and the ownership of data if mem-
bers leave, as well as roles of short-term members, such as graduate
students. These parameters should be formally specified and agreed
upon by all partners.
In addition to regular meetings, collaborators must routinely
send each other updates on recent data acquisitions, analyses, and
results. Some type of project schedule, such as a timeline or Gantt
chart, is also essential for coordinating and tracking progress on
different activities, such as fieldwork and data collection, running
analyses, developing and submitting grant proposals, and producing
publications. The updates and schedule should be accessible online
to all research project participants. Creating intermediate deadlines
for collaborators to come together to jointly review the progress to
date helps maintain the momentum of the project, and also ensures
that collaborators are still in agreement on the research questions
and overall research design. These reviews should also include oppor-
tunities to revise approaches and to modify research questions. Such
occasions often include some of the most productive discussions
about integration and interpretation.
Leadership is very important in the process of building trust.
The leadership role does not have to be the responsibility of a single
leader, but can be shared by several participants, with the key aim
of fostering an atmosphere conducive to trust. The Swedish sociolo-
gist Stefan Svallfors (2012) created the terms of ‘the good room’ and
‘the bad room’ in relation to seminars and collaborations. The good
room is characterised by positive relations, trust, safety, communica-
tion, and humility, while the bad room is characterised by insecurity,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
0
5
2

250 Jones et al.

feelings of uselessness, power games, and suppression techniques.


Active leadership can influence the path a group takes, and the room
the group builds.
In our experience, the most effective organisation of leadership
is through matrix management (Project Management Institute 2013)
or heterarchical structure (for a discussion of the definition and his-
tory of the term heterarchy, see Crumley 1995, 2005, 2007b). In these
models, the leadership role fluctuates among members according to
the expertise currently in demand. For example, a team member may
take the lead on a jointly written article to be submitted to a journal
within his/her particular discipline. For other activities, leadership
may fall to members according to personal aptitudes and predilec-
tions, such as knowledge of cutting-edge technology and computer
applications, a talent for scheduling and organising, or a personality
that finds it easy to interact with the public on behalf of the group.
Other time-consuming tasks, such as updating the schedule, taking
notes at group meetings, or posting material on shared webpages,
may be rotated to avoid overburdening any one member. This shar-
ing of leadership responsibilities keeps the group members on a more
equal footing, increases engagement in the collaborative effort, and
ensures that the talents and expertise of the group are used to the
fullest.
‘Followership’ is a recent concept that seeks to more fully
define the relationship between leaders and groups, and empha-
sises an active and dynamic role for team members not currently
in a leadership role. The idea is that, rather than passively follow-
ing the directions of the leaders, team members engage in proactive
and supportive followership, which mirrors the leadership role and
responsibilities. This approach is used successfully in settings with a
challenging environment and highly dynamic organisational context,
e.g. NASA training, and on the International Space Station, where
it is used by various astronauts and cosmonauts (Riggio, Chaleff, &
Lipman-Blumen 2008). There are always more followers than leaders
in any situation, and in a setting with shared or rotated leadership,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
1
5
2

How to Operationalise Collaborative Research 251

proactive, thoughtful, and dynamic followership becomes even more


important. Dynamic followership supports the leaders and actively
promotes the success of project structures and goals.

8.5 Language and Learning


Academic disciplines each have their own specific language, def-
inition of terms, and jargon, which impedes interdisciplinary com-
munication. Collaborators should use language that is as jargon free
as possible, in order to facilitate communication across disciplines
within the research group, and to make publications more access-
ible to wider audiences. One caveat is that the same word is often
defined differently in different disciplines or research specialties,
or may carry theoretical baggage in one discipline not apparent to
those with differing disciplinary perspectives, e.g. the term culture.
Interdisciplinary research groups, such as those working in histor-
ical ecology, need to agree on common definitions of terms useful
for their collaboration, such as landscape, economy, or adaptation.
In the same manner as collaborators deal with different terms used
by various disciplines, they also need to translate disciplinary inter-
pretive concepts and frameworks for each other. Ideally, such efforts
produce a kind of ‘epistemological Esperanto’1 that can be developed
by the group to facilitate interpretation of the factors or trends that
emerge from analyses performed across the different disciplinary
data sets (see also Chapter 6 about the advantage of different uses and
interpretations of terms by different disciplines). Examples of these
bridging concepts can be found in interdisciplinary frameworks, such
as historical ecology, where cross-disciplinary terms and concepts
have already been well established. For example, the term ecology
as developed within historical ecology includes the dialectical inter-
action of humans with their environment, in many senses negating
the duality of ‘natural’ and ‘cultural’ environments.

1
Term coined by Erik Gómez-Baggethun at the Historical Ecology Workshop: Is
There a Future for the Past? Meeting Challenges in the Research and Practice of
Historical Ecology, Odalgården, near Uppsala, Sweden, 16–18 April 2013.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
2
5

252 Jones et al.

A very important part of this process is learning. It is imperative


for an interdisciplinary project team to achieve at least a minimum
of understanding of terms and concepts between disciplines, and any
collaboration will benefit from initial efforts where participants learn
by listening to one another. The learning process takes time, and is
likely to result in a slower initial publication rate than for a group of
researchers working within the same discipline. Any interdiscipli-
nary project must plan and allocate resources for this learning phase.
While it is essential to develop a common understanding of terms
and concepts, it is also necessary for researchers to be able to access
the deep understanding and theoretical development inherent in the
different disciplines involved. Reading key texts from the different
disciplines followed by internal seminars on the meanings and inter-
pretations of those texts can be very helpful. Such discussions are
often the most rewarding parts of the interdisciplinary process and
may trigger new hypotheses or stimulate the development of new
theoretical approaches. Team members can at times feel threatened
on occasions when their frames of reference are challenged, which
again emphasises the importance of good leadership. In groups where
trust levels are high, more constructive and critical discussions are
possible, as participants do not feel personally challenged. The suc-
cess of such learning processes requires a certain amount of personal
courage since participants have to abandon their positions as experts
and instead become novices in relation to another field of knowledge.

8.6 Authorship and Publication


Due to the significant institutional barriers for the publication of col-
laborative, interdisciplinary research, a plan for equitable authorship
and publishing that reflects participants’ contributions and profes-
sional goals needs to be established at the beginning of each research
project. Even when an article or book is truly collaborative, with
equal effort from everyone participating in the writing and research,
it is impossible to give everyone equal credit, since the author listed
first in the publication receives the most recognition. In terms of

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
3
5
2

How to Operationalise Collaborative Research 253

career advancement in some disciplines or institutions, being listed


as the third or subsequent author awards no merit. Another diffi-
culty is the shortage of publishing outlets for truly interdisciplinary
work. The fact that different publishing venues do not carry the same
weight or status for the various disciplines also complicates mat-
ters. One way around this is to disseminate the research results in
interdisciplinary outlets as well as in single-disciplinary outlets that
match the different research specialties of the collaborators. In this
way, individual publication and professional advancement needs can
be met. A strategy that has worked for the authors entails collabo-
rators taking turns as lead author for interdisciplinary publications,
and in the case of more specialised publications, the collaborator
whose specialty matches the disciplinary publishing venue takes
the lead. Other publishing problems relate to the types of sources
that are acceptable as citations. In the natural sciences, sources in
English are preferred or even required. In the social sciences and the
humanities, however, a great deal of research is published in other
languages, usually that of the culture or society under study, or in the
researcher’s own language. This is due to disciplinary tradition and
also has certain advantages, such as using language as an analytical
tool, and using the language of the original sources, thereby avoiding
errors in the translation of cultural terms. These different publishing
conventions may set constraints on the publishing of interdiscipli-
nary research, especially in natural sciences publishing outlets, and
authors may have to present strong arguments for citing work writ-
ten in languages other than English.
An additional complication is related to the nature of inter-
disciplinary explanation. Publishing research that integrates several
disciplines introduces methods and types of data that may not be
familiar to all readers. In order to be transparent and comprehen-
sible to a wider audience, interdisciplinary publication generally
requires a more thorough explanation of methods and more detailed
descriptions of the data than publication within a single academic
discipline. This can be difficult as many journals limit the length

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
4
5
2

254 Jones et al.

of articles and also restrict the number and types of illustrations.


Although this situation is improving, some print publications still
require graphics in black and white or half-tones. This is often a real
impediment for the visual portrayal of complex and integrated data
sets. Online publications do not generally have such limitations on
colour or number of illustrations. One strategy for offsetting such dif-
ficulties is for collaborators to publish individual results in their own
standard single-discipline outlets where more extensive descriptions
of their data and methods can be developed (for examples, see Madry
2006; Jones 2009). This makes it possible to publish the integrated
group research in shorter, more focussed papers, while referring read-
ers to the separate single-discipline publications for more in-depth
elucidation of the data and methods involved.
A new and interesting collaborative forum has recently been cre-
ated through the websites www.academia.edu and www.researchgate
.net, which allows academics to share research papers and interests.
Academics around the world use it to ‘follow’ other researchers, post
their own papers, and track how many people have viewed their
papers. A recent feature at www.academia.edu is the ability to cre-
ate a ‘session’, which allows you to publish a paper or a document
of any type, and then invite other registered users to read it and post
their comments or questions. This feature allows each participant to
post general comments or to apply their comments, criticisms, and
ideas to a specific part of a text. At present, these sessions only last
for twenty-one days, but it is a very interesting new concept, which
allows a paper to be read and responded to by a large community of
scholars. This, and other, similar tools, has the potential to become
a highly useful means for receiving informal feedback and discussing
ideas with scholars outside the collaborative team before the formal
and rigorous peer-review process.
It is also critical for the research team members to negotiate and
understand each other’s professional objectives and requirements in
terms of publication, e.g. type of research dissemination (conference
paper, journal article, book, thesis, etc.), primary audience, optimal

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
5
2

How to Operationalise Collaborative Research 255

publishing time frame, and other considerations. As with the pri-


mary research objectives, e.g. questions to be answered, methods to
be tested, etc., these secondary professional objectives should also be
clearly laid out in the early stages of collaboration, and they may vary
considerably between researchers, especially at different stages of
their career. A junior faculty member may be under pressure to rap-
idly submit a number of articles to journals with a high impact fac-
tor, while senior tenured scholars may have less strong preferences
for prestigious journals, and are not always under the same pressure
to publish in the early phases of the collaboration. Differing work
schedules, such as year-round employment versus nine-month aca-
demic appointments, or varying holiday schedules and differing time
zones, may also create difficulties in setting a time frame for pub-
lication. These varying expectations, requirements, and objectives
can cause friction in the collaborative publication process. Thus, it is
important for all team members to communicate their expectations,
professional goals, work/travel schedules, and personal deadlines
early in the planning stages of the research, in order to minimise
group tension and prevent potential conflicts.

8.7 Funding
Funding opportunities for interdisciplinary research are fewer than
for single-discipline projects, despite the general acknowledgement
that interdisciplinary research has many advantages, and is impera-
tive for addressing complex research questions. The amount of fund-
ing required for a team of scholars rather than a sole researcher, and
for the longer periods required for first individual and then joint ana-
lysis and interpretation, can be quite substantial. Competition for
the rare major interdisciplinary funding opportunities is naturally
intense. Added to that is the fact that funders often have difficulties
in finding qualified reviewers for interdisciplinary project proposals,
resulting in the rejection of worthy proposals that are misunderstood.
The authors have found that a diverse funding strategy is success-
ful for sustaining long-term interdisciplinary team-based research,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
6
5
2

256 Jones et al.

consisting of both a major grant to cover the work of the whole team
for a set period of time, and smaller grants to fund specific aspects
of the research (e.g., particular analyses, such as radiocarbon dating,
pollen analysis, digitisation of spatial data for GIS, fieldwork, etc.).
Grants for the individual researchers on the team, pertaining to their
own single-disciplinary data collection and analyses – to be inte-
grated with the group effort at a later point – can also help keep the
collaborative research going. This type of approach requires flexibil-
ity as the pace and order of work will need adjusting according to the
fluctuating flow of funds. It also takes a great deal of planning ahead
in order to prioritise activities to best meet the various research goals
of the team.
Crowdsourced funding is a potential option for supporting inter-
disciplinary projects, and may be particularly useful for small pro-
jects in urgent need of funding. This method relies on donations from
a large number of people through web platforms and social networks,
and is used to support new products, artists, and scientists (Wheat
et al. 2013). Several specialised platforms currently target scientists
in addition to the popular all-inclusive platforms (Cadogan 2014).
Certain considerations need to be made before a crowd-funding
campaign is initiated. There may be policies in place requiring
researchers to gain approval from their employers for crowd-funded
projects, even if the projects in question use independent platforms.
Other issues include fees, e.g. to the crowd source platform, and tax
liabilities, which need to be built into the crowd-sourced budget.
Institutions may also require that overhead covering the use of their
infrastructure is included in the crowd-funded budget. Universities
and research institutions may complain that crowd funding does not
often contribute to their overhead costs, but the amounts raised by
crowd funding are usually fairly small and the loss to these institu-
tions is negligible. Also to be considered are overhead and staff time
associated with incentive products, usually related to the project,
developed for donors. This type of funding has additional potential
drawbacks that go beyond logistical and overhead issues. A particular

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
7
5
2

How to Operationalise Collaborative Research 257

issue may be that it does not carry the prestige granted by peer review,
and is therefore unlikely to be taken into consideration for career
advancement, e.g. in a tenure review (Kaplan 2013). The informal-
ity of the crowdsourced funding may also make it inapplicable when
applying for matching funds on other grants. A stronger criticism of
crowd funding is that it is ultimately unaccountable – funding does
not guarantee that the project will be carried out and there are no
mechanisms for returning the collected funds, although these con-
cerns are more and more being addressed by different platforms (Giles
2012; Weigmann 2013). Some universities have initiated their own
crowd-funding platforms that they encourage students and faculty
members to use, which allow them to collect overhead funds, keep
peer review controls in place, and demand a level of accountability.
No matter the source, the presentation of the interdisciplinary
research design is the key to successful funding. It must convey how
research questions and data types involved relate to the various dis-
ciplines, as well as to a general audience. It must also explain why
each disciplinary perspective in the collaboration is essential, and
show how the research will be successfully integrated, in terms of
both methods and theory. A good example of this is the application
for the HagmarksMISTRA project, which involved several years of
discussion with the funding body MISTRA, ‘The Swedish Foundation
for Strategic Environmental Research’. The aim was to find the best
possible systems solution for maintaining and developing biodiver-
sity and other cultural values based on land-use history, and finding
sustainable management practices beneficial for both biodiversity
and the farmer economy. The first funding application was rejected,
partly because the funders were unable to see sufficient integration
among the researchers involved. By the time funding was finally
granted, the project team had devoted a great deal of time and effort to
the presentation of the collaborative research design, which included
details of how the research would be integrated, joint research ques-
tions, the selection of common research sites, and setting up com-
munication with important stakeholders like the Federation of

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
8
5
2

258 Jones et al.

Swedish Farmers (sv. LRF) and Swedish World Wide Fund for Nature
(www.mistra.org). Often some degree of preliminary collaborative
research has to be conducted before applying for funding, so that
appropriate questions and strategies for integration can be worked
out. For long-term projects, it is essential that each team member has
a strong commitment to the common research objectives, in order to
keep the project moving forward during periods of meagre funding.

8.8 Integration of Data and Methods


When working with numerous and disparate types of data, it is cru-
cial to have some means for organising and visualising the data in
an integrated way. Various kinds of databases may be used, but the
authors have found GIS programmes to be particularly effective for
integrating data visually and spatially, as documents, photos, and
even audio-visual material can be linked to specific points in the
landscape through attribute tables and software such as eVis. For
historical ecology, the GIS may contain contemporary and histori-
cal map data (e.g. soils, roads, buildings, hydrology, land cover, and
elevations), remote sensing data (satellite images and aerial pho-
tographs), social network data (landownership, family relations,
demographic information, e.g. population, ethnicity, languages),
and contemporary and historical photographs showing changes
over time, etc. Thus, for historical ecology, a GIS is an efficient and
effective way of analysing and visualising the relationships between
humans and their environment (e.g. Swetnam, Allen, & Betancourt
1999; Donoghue & Moore 2003; Domaas 2007; Madry et al. 2011,
2015) GIS also provides a means to visually display data that span
different disciplines and to produce a common ‘view’ of a research
area, data, and the results of spatial analyses. Free and open-source
GIS software, such as QGIS, provide powerful GIS capability with
many advantages, including program translation into more than
forty languages.
Combining diverse types of data sets from different disciplines
requires the collaborators to be open to methods, analyses, and ways

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
9
5
2

How to Operationalise Collaborative Research 259

of presenting data that may not be the norm in their own disciplines.
Standards of ‘validity’ or testing vary among disciplines. Researchers
who predominantly use scientific methods are sometimes sceptical
about the validity of qualitative data and analyses. Yet it is long-
established that scientific methods, although perfect in their logic,
can never be perfectly applied (Kuhn 1996; Latour 1999, 2011). It
is impossible for humans to observe all factors relating to a set of
phenomena. This is especially difficult in the study of interaction
between humans and landscape, due to the sheer number of factors,
the technological limitations on our ability to observe, and/or socio-
cultural biases. Thus quantitative data are subject to selective pro-
cesses in their production and collection, and they always represent
a simplification of reality.
In disciplines such as history, sociology, and cultural anthro-
pology, only some of the data sets, such as grain prices and demo-
graphic trends, are quantifiable. Instead, the preponderance of data
(including oral history accounts and local knowledge collected
through ethnographic studies) is qualitative, related to specific con-
texts of time and place, and influenced by the unique perspectives
and experiences of individuals. Scientific methods are appropriate for
quantifiable data and repeatable phenomena. When this is not the
case, as with much of human history and society, the formal scien-
tific analytical structure is not always appropriate. Instead, evidence
is collected inductively with as much detail and coverage as possible.
Rather than relying on scientific proof, carefully drawn analogies
based on similar circumstances in other times and places can provide
means of interpretation and productive avenues for further investi-
gation. Although qualitative analyses lack the structure of deductive
methods, they are based on methods developed for source criticism
that rigorously examine how the data were produced, the purpose
for which they were produced, and the context in which they were
produced. Validity and accuracy for both qualitative and quantitative
analyses are improved by increasing the quantity and variety of data
sources related to a phenomenon. This is one of the great advantages

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
0
6
2

260 Jones et al.

of interdisciplinary work, that more factors relating to a specific phe-


nomenon can be examined.
An example of how the different types of data enhance each
other, and how data from one set can be used to refine or correct data
in other data sets, can be drawn from the authors’ experience of stud-
ying the history of ponds and water use in southern Burgundy (Madry
et al. 2011, 2013, 2015). In the course of researching a mill and pond
at Bauzot (Issy l’Évêque, Saône-et-Loire, France), gaps and errors in
the existing maps were identified and corrected using information
collected from other data sets. Ten maps, dating from 1759 to 2003,
were integrated into the GIS. The large mill pond was depicted on
the maps from 1759 to 1895, but not on those dating between 1912
and 1983 (Figure 8.4). At some point, then, between 1895 and 1912,
the man-made pond and mill had disappeared. Oral history accounts
indicated that they had been destroyed by a flood in the early twen-
tieth century. A postcard of the intact pond and mill dated to 1906
(Figure 8.5), and that, together with agricultural records, enabled us
to pinpoint the flood to sometime in 1907. However, the pond again
appeared on the map from 2003, leading us to assume that it had been
reconstructed in its former location after almost 100 years. This idea
was initially supported by research showing that a number of new
ponds were built in the area between 1983 and 2003. However, as
our study also included an oral history component, we interviewed
the proprietor of the Bauzot pond. He had owned the land since
2003 and informed us that no pond existed that at that time or in
the immediately preceding or subsequent years. The depiction of a
pond on the 2003 map was a cartographic error that we were able to
detect through the collection of oral history data and further verified
through other documentary sources.
The discrepancy between the map and the owner’s insistence
that no pond existed in 2003 is due to the fact that virtually all ponds
in this area are man-made by damming small streams. The 2003 map
was based on aerial photos taken in 2002 at a time of year when
weather conditions may have created pooled water in the area of the

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
1
6
2

How to Operationalise Collaborative Research 261

Figure 8.4 Maps of mill and pond area at Bauzot, France, 1838 (top
left), 1912 (top right), 1983 (bottom left), 2002 (bottom right). The pond
area has been marked with black; in maps where pond is missing, the
area is encircled.
Maps: Issy l’Évêque Tableau d’Assemblage, 1838 (Courtesy the Archives
Départementales de Saône-et-Loire). Modified from the1912 État Major.
Modified from the 1983 and 2002 1:25000 Luzy Topographic maps,
courtesy IGN (Institut Géographique National de France).

former man-made pond. Since this area was dry for most of the year,
and the owner did not repair or maintain the dam that had originally
created the pond, he did not recognise this area as a ‘pond’. Knowledge
of how the maps were created and of how local people defined ponds
proved essential for refining the data in our study.
Although simple comparisons among data sets as described
in this example, are fairly straightforward, correlating entire sets of
quantitative and qualitative data for complex analyses is never a sim-
ple process (Yang et al. 2014; Lennartsson et al. 2015). The social

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
2
6

262 Jones et al.

Figure 8.5 Postcard of Bauzot Mill building and pond circa 1906.
(Courtesy Archives Départementales de Saône-et-Loire, Mâcon.)

sciences have a long history of integrating qualitative data with


quantitative trends, and can therefore provide useful models for
interdisciplinary research (Barber & Berdan 1998; Sampson & Laub
1998; Burke 2008; Austin et al. 2015; Creswell 2015). Even within
this framework, exploration of different ways of comparing data,
integrating analyses, and interpreting the results is required. In most
cases, the integration of different data sets creates new insights, helps
identify which in-depth analyses will be suitable, and typically leads
to additional research questions.
In order to facilitate validation and comparative analyses, it is
important that studies can be replicated, as this allows researchers to
revisit them at some future date in order to incorporate new data, to
expand the study, or to apply new interpretive methods or different
theoretical frameworks. Therefore, in the early phases of a project,
collaborators must agree on secure and robust methods of handling
all types of data, and decisions on how to proceed with the analyses
must be documented.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
3
6
2

How to Operationalise Collaborative Research 263

Ideally, the spatial-temporal scale of each data set should be


roughly comparable in order to effectively integrate data, run analy-
ses, and interpret the results. However, a certain degree of historical-
ecological analysis can be carried out at different analytical, spatial,
and temporal scales, as the factors that influence a particular land-
scape or ecosystem will reflect a series of overlapping and nested
hierarchies of influence and interrelationship, both spatial and tem-
poral. Therefore, differing data sets can be at least partially integrated
in this nested way.
Identifying such nested relationships is important, but it is
almost always fruitful to focus on the impact of broad-scale factors
(such as global climate change, national/international changes in eco-
nomic regulations, or regional transportation network development)
at the local scale. By applying a great degree of temporal precision, it
is possible to study how these constraints/facilitators influence local
and relatively short-term processes, reactions, and decisions. It has
proven rewarding for the authors to combine data at very detailed
analytical levels, where the mechanisms and links between humans
and environment are often clearer. In this way, misinterpretations as
a result of the generalisations which are required when working at
regional levels or in broader time scales may be avoided (Madry et al.
2011, 2015). Information about the environmental requirements of
a demanding species can, for example, reveal a great deal about past
and present landscapes, if combined with detailed land-use informa-
tion on the local scale (Dahlström et al. 2013). Such detailed indi-
vidual studies can then be linked over larger geographical areas and
longer time spans in order to provide a bigger picture of temporal and
spatial variation.

8.9 Data Management


Managing the large amounts of data accumulated in collaborative,
long-term projects can be a logistical nightmare. It is important that
collaborators have easy access to one another’s data and analyses,
in order to facilitate thinking about integration, relationships, and

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
4
6
2

264 Jones et al.

influence/causation. This is very difficult in practice, but can be done


with proper planning and organisation. Members of a project team
often belong to different, geographically separated institutions, and
membership may fluctuate as new collaborators join and others leave
or retire, etc. It is therefore vital to archive data properly, and plan
and organise for this in the early stages of the research project. Some
type of online platform for archiving the project data, results, current
work, and publications is essential if collaborators are to have easy
access. Such archives also provide backups in case of accidental loss
of data and results. Although there are advantages to online archives,
actually working online with large data files can sometimes be dif-
ficult due to bandwidth limitations. Consequently, it is necessary to
have a process for creating systematic versions of data files to avoid
duplication of effort and loss of content, as files are downloaded,
revised, and uploaded again. Standard practices, such as adding dates
to filenames of documents and graphics, facilitate control over dif-
ferent versions. Standardised filenames further assist collaborators
in locating data. Creating an index of group holdings and metadata
descriptions for files is essential for large and long-term projects.
Managing version control during the collaborative writing pro-
cess is also critical when working with multiple authors. Various
tools, such as ‘track changes’, in word processing software allow
users to monitor the evolution of the text, as well as provide for the
insertion of author comments on points that need discussion and
a collective decision. Common internet spaces, e.g. wiki pages, are
requisite for large and geographically separated collaborative groups.
These sites facilitate collaborative writing, enable online group con-
versations, and provide shared electronic calendars for long-term
planning, tracking work, and displaying individual availability.
Digital archiving is a rapidly evolving technology. Powerful,
open-source tools allow the creation, search, and display of vast
archives of disparate data, including photographs, records, sound
recordings, video, digital documents, and GIS data. Resources of this
type have often been a final step for saving data at the end of projects,

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
5
6
2

How to Operationalise Collaborative Research 265

or when researchers retire, but have recently also been used during
active research. The Odum Institute’s Dataverse is an example of a
repository that is useful throughout the life of projects, as it allows
for specification of access that may change over time. It is recom-
mended that such tools are incorporated at the beginning of a pro-
ject, and adding a digital archiving specialist to the group will greatly
assist in making that possible. Complex interdisciplinary projects
benefit from employing a librarian, skilled in digital archiving and
data management, in order to assist with designing digital data man-
agement systems (including metadata and appropriate file standards).
This approach is becoming more common in a wide variety of aca-
demic endeavours (Kesselman & Watstein 2009).

8.10 Conclusions and Collaboration Checklist


It is always more difficult, slower, and less efficient to work collab-
oratively (‘by committee’) than it is to conduct individual research.
Reaching consensus is difficult, both in day-to-day operations and for
long-term goals. Collaborative research, moreover, does not always
bring the same career rewards and prestige as individual accomplish-
ments. Despite all this, there are many epistemological advantages of
collaborative, interdisciplinary work which are becoming indispens-
able to adequately address the types of interrelated questions posed
by historical ecology. A bonus is the intellectual stimulation and cre-
ativity produced by a group of diverse people addressing common,
complex issues. The new insights and expanded perspectives that can
be gained are rewards in themselves, both on a personal level and for
the advancement of scholarship as a whole.
This chapter has provided an outline of practical considerations
for making interdisciplinary team research more harmonious, effi-
cient, and effective. Our aim is to provide researchers with a ‘check-
list’ for matters that require explicit attention in any collaborative,
interdisciplinary effort. This list is not meant to represent a sequen-
tial or hierarchical ordering (Figure 8.6), because the circumstances
in every team will differ, and thus the order of addressing issues will

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
6
2

266 Jones et al.

Figure 8.6 Schematic of issues to be formally addressed in


collaborative research.

vary, e.g. some groups are brought together by existing funding, while
others will work together for some time before applying for major
funding. Most groups will develop their own solutions and methods,
but we have found – sometimes the hard way – that these are areas
that cannot be ignored or left to evolve haphazardly if a project is to
succeed.

8.10.1 Collaborative Research Checklist

• At the outset, dedicate time for mutual learning and the development of
a common set of terms and concepts through presentation of individual
research, reading of key background literature, and group discussion.
• Develop overarching research questions that are as relevant to the
individual disciplines involved in the collaboration as they are to the
integrated research effort.
• Involve local inhabitants and other non-academic stakeholders with
traditional or local knowledge as collaborators, and include the research
questions or objectives that are meaningful to them.
• Use practices that promote trust and open communication, and establish
rules for group membership, data sharing, and ownership of results.
• Set up a framework for effective leadership, followership, and the
equitable sharing of responsibilities that may include a rotation of duties

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
7
6
2

How to Operationalise Collaborative Research 267

and/or a heterarchical organisation of shifting leadership/followership


roles, according to context.
• Develop strategies for equitable publication in both interdisciplinary
and intra-disciplinary venues, as well as publishing and disseminating
information in formats accessible to lay stakeholders and to the general
public.
• Develop and use a variety of strategies for funding of long-term
collaborative efforts, and prepare for the funding process by carrying
out some initial, exploratory collaborative research and experiments,
including integrating data sets in order to refine the research questions
and strengthen the proposed research design.
• Experiment with different models and approaches for integrating data
and methods, taking into account: (1) the different physical formats of
data; (2) differing standards of validation for different types of data; (3) the
comparison, correlation, and integration of quantitative and qualitative
data; and (4) issues of scale; and always document the decisions made on
the selected practices for validating data sets and for choosing modes of
analysis, to ensure that the study can be replicated, expanded, or revisited.
• Create an initial plan for data management, including shared and
remote access, version control, and backup protocols, and, if possible,
include collaborators with archiving/data management expertise in the
research team.

In order to succeed, interdisciplinary research has to be analytically


integrative and intrinsically collaborative. It is inadequate to say that
a project is interdisciplinary simply because it has a broad scope or
vision (Thompson-Klein 2012). Even though shared methods, research
questions, and concepts are key stepping stones for interdisciplinary
activities, we would like to propose that a longer-term goal should
be to go beyond interdisciplinary research, in pursuit of Jean Piaget’s
transdisciplinary goals of purposefully transgressing the conventions
and boundaries of established disciplines (1972). Transdisciplinary
initiatives not only bridge the divide between academic knowledge
and lay knowledge, but generate new scaffolding paradigms that may
transform the focal points and scopes of disciplines. Such new para-
digms have the capacity to transcend the perspectives of traditional

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
8
6
2

268 Jones et al.

academic disciplines as well as the boundaries between professional


and lay knowledge in order to create new, holistic approaches for
addressing socio-ecological complexity. By examining the practical
considerations of collaborative interaction and research, we hope to
help free groups to focus on these theoretical goals and to pave the
way for genuinely transdisciplinary achievement.

References
Austin, Z., McVittie, A., McCracken, D., Moxey, A., Moran, D., & White, P.
C. L. (2015). Integrating quantitative and qualitative data in assessing the
cost-effectiveness of biodiversity conservation programmes. Biodiversity and
Conservation, 24(6), 1359–75. doi: 10.1007/s10531-015-0861-4.
Baleé, W., ed. (1998). Advances in Historical Ecology. New York: Columbia
University Press.
Barber, R. J. & Berdan, F. F. (1998). The Emperor’s Mirror: Understanding Cultures
through Primary Sources. Tucson: University of Arizona Press.
Boix-Mansilla, V. (2010). Learning to synthesize: the development of interdisciplin-
ary understanding. In R. Frodeman, J. T. Klein, & C. Mitcham, eds., The Oxford
Handbook of Interdisciplinarity. Oxford: Oxford University Press, pp. 288–306.
Burke, P. (2008). What Is Cultural History? 2nd ed. Cambridge: Polity Press.
Cadogan, D. (2014). Funding for research? Look to the crowd. College & Research
Libraries Newsvol, 75(5), 268–71.
Chettiparamb, A. (2007). Interdisciplinarity: A Literature Review. Southampton,
UK: University of Southampton. The Interdisciplinary Teaching and Learning
Group, Subject Centre for Languages, Linguistics and Area Studies, School of
Humanities.
Choi, B. C. & Pac, A. W. (2006). Multidisciplinarity, interdisciplinarity and trans-
disciplinarity in health research, services, education and policy: 1. Definitions,
objectives, and evidence of effectiveness. Clinical and Investigative Medicine,
29(6), 351–64.
Creswell, J. W. (2015). A Concise Introduction to Mixed Methods Research. Los
Angeles: Sage Publications.
Crumley, C. L., ed. (1994). Historical Ecology: Cultural Knowledge and Changing
Landscapes. Santa Fe, NM: School of American Research Press.
Crumley, C. L. (1995). Heterarchy and the analysis of complex societies.
Archaeological Papers of the American Anthropological Association, 6, 1–5.
doi: 10.1525/ap3a.1995.6.1.1

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
9
6
2

How to Operationalise Collaborative Research 269

Crumley, C. L. (2005). Remember how to organize: heterarchy across the disci-


plines. In C. S. Beekman & W. S. Baden, eds., Nonlinear Models for Archaeology
and Anthropology, Aldershot, UK: Ashgate Press, pp. 35–50.
Crumley, C. L. (2007a). Historical ecology: integrated thinking at multiple tem-
poral and spatial scales. In A. Hornborg & C. L. Crumley. eds., The World
System and the Earth System: Global Socio-environmental Change and
Sustainability since the Neolithic. Walnut Creek, CA: Left Coast Press,
pp. 15–28.
Crumley, C. L. (2007b). Heterarchy. In W. A. Darity, ed., International Encyclopedia
of the Social Sciences, 2nd ed. Detroit, MI: Macmillan Reference USA, pp. 468–9.
Crumley, C. L. & Marquardt, W., eds. (1987). Regional Dynamics: Burgundian
Landscapes in Historical Perspective. New York: Academic Press.
Dahlström, A., Lennartsson, T., & Iuga, A. M. (2013). Managing biodiversity rich
hay meadows in the EU: a comparison of Swedish and Romanian grasslands.
Environmental Conservation, 40(2), 194–205.
Domaas, S. T. (2007). The reconstruction of past patterns of tilled fields from his-
torical cadastral maps using GIS. Landscape Research, 32(1), 23–43.
Donoghue, M. J. & Moore, B. R. (2003). Toward an integrative historical biogeog-
raphy. Integrative and Comparative Biology, 43, 261–70.
Frodeman, R., Klein, J. T., & Mitcham, C. (2012). The Oxford Handbook of
Interdisciplinarity, Reprint edition. Oxford: Oxford University Press.
Giles, J. (2012). Like it? Pay for it: with conventional sources of money drying up,
some scientists are turning to crowd-funding. Nature, 481, 252–3.
Jaarsma, S. (2002). Handle with Care: Ownership and Control of Ethnographic
Materials. Pittsburgh, PA: University of Pittsburgh Press.
Jones, E A. (2009). Multi-temporal landscape history in Burgundy: an innovative
application of genealogy software. In B. Frischer, J. W. Crawford, & D. Koller,
eds., Making History Interactive. Computer Applications and Quantitative
Methods in Archaeology (CAA). Proceedings of the 37th Annual International
Conference on Computer Applications to Archaeology. Oxford: British
Archaeological Reports.
Kaplan, K. K. (2013). Crowd-funding: cash on demand. Nature, 497, 147–9.
doi: 10.1038/nj7447-147a
Kesselman, M. A. & Watstein, S. B. (2009). Creating opportunities: embedded
librarians. Journal of Library Administration, 49(4), 383–400.
Kuhn, T. S. (1996). The Structure of Scientific Revolutions, 3rd ed. Chicago:
University of Chicago Press.
Latour, B. (1999). Pandora’s Hope: Essays on the Reality of Science Studies.
Cambridge, MA: Harvard University Press.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
0
7
2

270 Jones et al.

Latour, B. (2011). From multiculturalism to multinaturalism: what rules of method


for the new socio-scientific experiments? Nature and Culture, 6(1), 1–17.
Lennartsson, T., Westin, A., Erikson, M., Flygare, I. A., Isacson, M., & Morell, M.
(2015). Between nature and society: interpretation of an early 19th century
farmer’s diary. Agricultural History Review, 63(2), 265–85.
Madison, D. S. (2012). Critical Ethnography: Method, Ethics, and Performance,
2nd ed. Thousand Oaks, CA: Sage Publications.
Madry, S. (2006). Hic sunt dracones (Here be dragons): The integration of his-
torical cartographic data into geographic information systems. In S. Archer
& K. Bartoy, eds., Between Dirt and Discussion: Methods, Materials, and
Interpretation in Historical Archaeology. New York: Springer, pp 33–60.
Madry, S., Jones, E. A., Murray, S., & Tickner, A. (2011). Une Micro-Histoire De
La Terre Et De L’utilisation Des Ressources: L’intégration des SIG-H (Systèmes
d’information géographique historiques) et des données qui y sont liées en
Bourgogne du sud. Le Monde des Cartes, Rapport Cartographique National
revue de Comité Français de Cartographie, 208, 75–94.
Madry, S., Jones, E. A., & Tickner, A. (2013). Interdisciplinary History of Rural
Water and Land Use in Southern Burgundy, France. Technical report. www
.doaks.org/ research/ garden- landscape/ garden- and- landscape- project- grant-
reports/madry-report.
Madry, S., Jones, E. A., Tickner, A. Murry, S., & Misner, T. (2015). Water and
landscape dynamics in southern Burgundy: two and a half centuries of water
management in an agricultural landscape. Water History, 7(3), 301–35.
doi: 10.1007/s12685-015-0132-z.
Moen, J. (2015). On viewing your belly button or the world. In E. Mineur &
B. Myrman, eds., All of Science! 15 Researchers on Integrated Research.
Stockholm: The Swedish Research Council, pp. 34–9.
Okely, J. (2012). Anthropological Practice: Fieldwork and the Ethnographic
Method. London: Bloomsbury Publishing.
Olsson, R. (2008). Mångfaldsmarker: Naturbetesmarker en värdefull resurs.
Uppsala: Centrum för biologisk mångfald.
O’Reilly, K. (2012). Ethnographic Methods, 2nd ed. London, New York: Routledge.
Piajet, J. (1972). The epistemology of interdisciplinary relationships. In L.
Apostel, ed., Interdisciplinarity: The Problems of Teaching and Research in
Universities. Paris: Organization for Economic Cooperation and Development,
pp. 27–39.
Project Management Institute. (2013). A Guide to the Project Management Body
of Knowledge (PMBOK Guide), 5th ed. Project Management Institute.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
1
7
2

How to Operationalise Collaborative Research 271

Riggio, R. E., Chaleff, I., & Lipman-Blumen, J., eds. (2008). The Art of
Followership: How Great Followers Create Great Leaders and Organizations.
San Francisco: Jossey-Bass Press.
Sampson, R. J. & Laub, J. H. (1998). Integrating quantitative and qualitative data. In
J. Z. Giele & G. H. Elder Jr., eds., Methods of Life Course Research: Qualitative
and Quantitative Approaches. Thousand Oaks, CA: Sage Publications, pp.
213–30.
Schensul, J. J. & Lecompte, M. D. (2013). Essential Ethnographic Methods: A Mixed
Methods Approach, 2nd ed. Lanham, MA: AltaMira Press.
Strang, V. (2009). Integrating the social and natural sciences in environmental
research: a discussion paper. Environment, Development and Sustainability,
11, 1–18.
Svallfors, S. (2012). Kunskapens människa: Om kroppen, kollektivet och kunska-
pspolitiken. Stockholm: Santerus.
Swetnam, T. W., Allen, C. D., & Betancourt, J. L. (1999). Applied historical ecol-
ogy: using the past to manage for the future. Ecological Applications, 9(4),
1189–1206.
Thompson-Klein, J. (2012). A taxonomy of interdisciplinarity. In R. Frodeman, J.
T. Klein, & C. Mitcham eds., The Oxford Handbook of Interdisciplinarity.
Reprint edition. Oxford: Oxford University Press, pp. 15–30.
Weigmann, K. (2013). Tapping the crowds for research funding. EMBO Reports, 14,
1043–6. doi: 10.1038/embor.2013.180.
Wheat, R. E., Wang, W., Byrnes, J. E., & Ranghanathan, J. (2013) Raising money
for scientific research through crowdfunding. Trends in Ecology & Evolution,
28, 73–4.
Whyte, K. P. (2013). On the role of traditional ecological knowledge as a collabora-
tive concept: a philosophical study. Ecological Processes, 2(7).
Yang, Y., Zhang, S., Yang, J., Chang, L., Bu, K., & Xing, X. (2014). A review of his-
torical reconstruction methods of land use/land cover. Journal of Geographical
Sciences, 24(4), 746–66.

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
2
7

Downloaded from https://www.cambridge.org/core. University of New England, on 23 Feb 2018 at 16:51:50, subject to the Cambridge Core
terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.008
3
7
2

Part III Moving Forward

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:48:19, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
4
7
2

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:48:19, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
5
7
2

9 Historical Ecology in
Theory and Practice: Editors’
Reflections
Tommy Lennartsson, Anna Westin, and
Carole L. Crumley

Taking part in discussions and listening to research experiences with


a group of open-minded and interdisciplinary researchers has given
us much to think about. Some of our colleagues engage with con-
cepts, theoretical tools, and research methodology. Others are of a
more applied nature and are interested in how historical ecology can
be developed to better understand the present and shape the future.
In recognition that both are central to historical ecology’s research
programme, we offer some of our own reflections on the group’s
collective work.

9.1 Returning to Causation and Complexity


Chapter 3 underscores the importance of linear time in assessing the
role of climate in societies, and evokes for the reader the temporal
conundrum that was introduced in Chapter 2. It showcases careful
temporal work at spatial scales less than global (climate evidence –
see Fact Box 3.1 in Chapter 3 – is collected and combined for locali-
ties, landscapes, and regions), and gives ample reason for expressing
change in linear temporal form. Given what is often temporally
precise evidence for the unfolding of disasters, it would seem that
clear causal links can be made between adverse weather or climate
and its direct effects on human health and well-being.
But what can such work say about causality? We have learned
that some factors that have reduced resilience in the past include a mis-
match of population size with resources or technical constraints, wide-
spread poverty, social inequality, underdeveloped trade infrastructure,

275

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
6
7
2

276 Lennartsson et al.

and lack of agricultural and economic diversity. In Chapter 3, how-


ever, it is emphasised that these links, while clear in particular cases,
are not sufficient to support prediction. Instead, to measure the vul-
nerability or robustness of past or contemporary societies – and to plan
for the future – requires knowledge of the social, political, economic,
and other contexts in which the environment influences humans, and
in which catastrophic events occasionally occur.
To manage risk is to manage a complex system, and so-called
environmental disasters must be holistically treated. Hoffman and
Oliver-Smith (2002) term societies that are unprepared ‘cultures of
catastrophe’. By this they mean that disasters previously seen as acts
of God or of Nature (thus out of human control) should instead be
seen as a complex system in which physical, biological, and social
factors together shape both a society’s readiness and its response to
catastrophe.
Aligning with disaster management professionals, a meta-
theoretical and holistic approach to risk management for historical
ecology combines linear with holistic/complex understandings of
time. Article 17 of the United Nations’ Sendai Framework for Disaster
Risk Reduction aims to ‘Prevent new and reduce existing disaster risk
through the implementation of integrated and inclusive economic,
structural, legal, social, health, cultural, educational, environmental,
technological, political and institutional measures that prevent and
reduce hazard exposure and vulnerability to disaster, increase pre-
paredness for response and recovery, and thus strengthen resilience’
(United Nations 2015). Thus while adverse weather and climate can
wreak havoc on human societies, it should be considered only one
of several interacting factors that together reduce societal resilience.

9.2 Can Insights from Historical Ecology


Travel through Time and Space?
9.2.1 Space Travel
Conservation in practice needs knowledge about, for example, trad-
itional land use techniques, the handling of natural dynamics and

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
7
2

Historical Ecology in Theory and Practice 277

variation, and the processing and use of natural products. Regions


and communities in which local ecological knowledge is still diverse
and in use are more and more in focus, for example as ‘biocultural
refugia’ (Barthel, Crumley, & Svedin 2013a, 2013b, see Chapter 4).
Chapter 4 discusses how new historical-ecological insights can be
made in one place (Sweden) through the comparison with another
place (Romania). When it comes to biodiversity and agrarian land
use, the similarities are striking between current Romania and the
past of many Western European countries. These similarities offer
an important source of information for conservation biologists
who struggle with designing conservation activities for threatened
habitats that are formed by bygone land use. We will return to this
problem in a later section of this chapter. Many are the biologists
who have come to Transylvania and returned to their home coun-
tries inspired by the traditional land-use practices and rich flora and
with new ideas about suitable management regimes of rare habitats
(Helldin & Lennartsson 2007). Phrased somewhat differently, places
which are rich in biocultural diversity (Chapter 7) can serve as, or
at least contribute to, baselines, or reference ecosystems (Egan &
Howell 2001; Gavin et al. 2015), for nature conservation elsewhere
(Chapter 5).
How to use local biocultural knowledge in other places is a
potentially fruitful research field, and so far there is not much experi-
ence to lean on. The transport of biocultural knowledge must be
treated with caution. How, and under what circumstances, can the
understanding of human-environment links be ‘translated’ before
being used in other places, which most likely have different nat-
ural and societal contexts? One way forward will be to focus on the
detailed links between practices and ecology. A basic assumption is
that species react similarly to the same type of disturbance no matter
where it happens, such as a plant being cut at a specific length in a
specific time in its life cycle. The same disturbance (land-use prac-
tice component) can be the result of various social-economic con-
texts. These are relevant for understanding, e.g. the links between a

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
8
7
2

278 Lennartsson et al.

community’s knowledge and its land-use practices. But for a plant,


it makes little or no difference who was holding the cutting tool,
and what was the social context of the work. At this level of detail,
it is possible to translate knowledge without making mistakes. The
example of cutting of plants has highly relevant practical applica-
tions, namely, finding suitable mowing dates for vulnerable species
and meadow habitats (Eriksson et al. 2015).

9.2.2 Time Travel


What if the combined study of a region’s historic climate and socio-
economic conditions could be used more extensively than is now the
case to produce scenarios for its future? With detailed evidence of
historic change for many regions, an interesting opportunity presents
itself. It is possible to trace changes in precipitation, temperature,
and seasonality, as well as to assess the intensity, duration, and fre-
quency (IDF) of events and periods. Proxy data would need to be used
(see e.g. Ljungqvist et al. 2016), including careful reassessment of evi-
dence from many sources, such as dendroclimatology, documents,
and geomorphology. In addition to more typical changes, long-term
proxy data offer an overview of notable events and periods which are
well beyond regional norms, termed excursions. While clarification
of what is meant by the term excursion in specific regional contexts
would need attention, the possibility exists of finding conditions that
match climate model–based regional scenarios (currently associated
with large uncertainties) by introducing readily available regional
data such as terrain, water management, diaries, and other means
of assessing human preparedness and response. All regions have his-
toric periods of dramatic hydroclimate excursions and periods when
the frequency of events (e.g. storms, hurricanes, droughts) is high. In
order to understand the effects of climate on societies, such condi-
tions, as well as contemporary events, could be studied by combin-
ing climate data with information about the economic, social, and
ecological conditions that exacerbate difficulties, and about physical
effects on hydrology, soils, etc.

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
9
7
2

Historical Ecology in Theory and Practice 279

Thus it may be possible to work both back in time and forward


into the future (termed, respectively, backcasting and forecasting) to
augment models of future conditions. Not the least, this approach
can show how societies’ vulnerability and preparedness to climate
events may depend on resilient use of ecosystems.
Four decades of collaborative historical ecological research in
Burgundy (France) has given insight into centuries of shifting cli-
mate, terrain, hydrology, husbandry, and social organisation (e.g.
Crumley & Marquardt 1987; Gunn & Crumley 1991; Gunn et al.
2004; Jones 2006; Madry et al. 2015). In part due to the region’s his-
toric prominence in viticulture, there is keen and enduring interest
in Burgundy’s weather and climate (e.g. Le Roy Ladurie et al. 2006).
Recently, periods of lengthy regional drought have increased; among
them are three truly disastrous droughts in 1976, 2003, and 2015.
The 2003 drought was especially severe in Burgundy; animals died
and prices plunged, pastureland burned, and persistent temperatures
in the high 40s °C took the lives of 20,000 people in France alone. In
general these droughts were triggered by a huge low-pressure system,
stalled for six months or more in the Atlantic, that turns rain away
from the Atlantic façade from Spain to the United Kingdom (Crumley
1993). This high-summer Mediterranean pattern can now begin in
the winter and last more than six months. This is an example of an
excursion which could be a harbinger for a warming world, and of a
vulnerable region where researchers have already studied the historic
and contemporary agro-economy and landscape-scale environmental
and human responses to historic and contemporary excursions.

9.3 Borrowing Concepts


Definitions, terms, and concepts have a long history of being passed
back and forth in the environmental and social sciences, with vary-
ing results. The impetus is understandable – to study and communi-
cate about a complex world – but the dangers are many. Examination
inevitably reveals the weaknesses of any concept, not least when a
concept coined in one discipline is scrutinised by another. Not less

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
0
8
2

280 Lennartsson et al.

difficult is making sense of contemporary differences of opinion in


one’s own field of study. This leaves the intellectual history of dis-
ciplines littered with battered or abandoned efforts and, ironically,
offers insight into a recurrent tendency to ‘borrow back’, forming a
sort of braided stream of ideas (Ellen 1982; Kellert 2008). This is not
entirely due to differences between disciplines that share a concept
or term; as time goes on, the practice of the scientific method and
within-discipline critique propel the definition, term, or concept into
new territory.

9.3.1 Going beyond Definitions


It is especially interesting to apply concepts in a way that reveals
causal links between humans and their environments, as discussed
in several chapters of this book. Causal links represent the character
of human-environment interactions, which shape both. Interactions
are not the same as correlations, as discussed for the concept of bio-
cultural diversity in Chapter 7. We see that straightforward corre-
lation estimates, such as those combining number of species in a
region with the number of cultural expressions (e.g. languages), do
not necessarily reflect the number of actual interactions. Elaborate
estimates and historical-ecological knowledge are needed for describ-
ing the diversity of contact points at which cultural and biological
diversity meet, interact, and influence each other.
In Chapter 6, the concept of domestication (of animals and
plants) is seen to have shifted from a distinct change in state, being part
of an ‘agricultural revolution’, to domestication as a process, drawn
out over several thousand years and gradual in terms of morphologi-
cal change of the domesticated species. Nevertheless, this gradual
process may contain steps, such as obtaining the capacity for culti-
vating certain plants (in contrast with collecting them in the wild),
and for keeping and herding animals (in contrast to hunting them).
Viewed as a revolution from the perspective of subsistence, these
changes transformed annual scheduling, settlement form and perma-
nence, technology, religion, human health, and social structure, to

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
1
8
2

Historical Ecology in Theory and Practice 281

name only a few. Furthermore, analysing causal human-environment


relationships that are involved in domestication can show that not
all types of domestication have been equally gradual. Milk, in con-
trast to meat, can only be collected if the animals are under control
and tame. Many cereals and vegetables cannot be cultivated in cold
climates without collecting the seeds or other propagation units and
storing them under certain conditions over the winter. In such cases,
we see rather sharp shifts from non-domesticated to domesticated.
An example of a more gradual domestication is that before grain
plants became actively cultivated by manual sowing on prepared
fields, some may have been deliberately favoured by regular selec-
tion of more resistant seeds and the construction of suitable niches
for these specific plants. Another example is the semi-domesticated
reindeer, discussed in Chapter 6. A discussion about how we use the
term domestication can thus reveal domestication processes involv-
ing different interrelated responses among human societies, environ-
mental conditions, and domesticated species.

9.3.2 Taking Advantage of Differences


Any interdisciplinary collaboration must discuss the use of terms
and concepts, some of which have different meanings in different
contexts and disciplines. Collaboration may lead to an agreement
of the use of common terms, in Chapter 8 called ‘epistemological
Esperanto’. It is an important part of collaboration to understand
each other, but it is also important not to lose the complexity of the
terms as they are employed in different disciplines. In the process of
finding ‘common ground’, there may be a risk that the common use
of a concept becomes poorer in meaning and perhaps less useful as
an analytical tool.
It may be discouraging to read about difficulties and traps in
the swamp of definitions, but the comparison of how concepts are
used in different disciplines can also uncover previously undetected
human-nature relationships. In attempting to harmonise termin-
ology one must first understand how the terms have been used by

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
2
8

282 Lennartsson et al.

others. After that it is possible to relate more freely to the terms and
it will be easier to define how the terms are used in an interdisciplin-
ary group or publication.
Development of interdisciplinary grammars and the identifica-
tion concepts which are shared or contested are fundamental to the
practice of historical ecology (Newell et al. 2005). A good example
is the term landscape (discussed in Chapter 6), a unit of analysis in
many academic disciplines (archaeology, architecture, art, ecology,
geography, geomorphology, regional planning) and also familiar to the
general public. Thus landscapes are read and interpreted by everyone
as likely to promote lively discussion in a gathering of citizens as in a
group of scholars from various disciplines. Researchers and the pub-
lic benefit from the understanding of terms’ different perspectives.

9.4 Interdisciplinary and Transdisciplinary


Collaboration
Human-environment relations emphasise the need for deep interdis-
ciplinary methods for collaboration, as discussed in Chapter 8, which
is a question of both what we study and how we do it. In order to real-
ise the full potential of the historical-ecological approach, it should
go beyond applying a natural-historical perspective on culture and
a cultural-historical perspective on nature, and beyond attempts to
find ecological explanations for cultural expressions and vice versa.
An important task for historical-ecological analyses is instead to
discover the explanatory potential that may lie in the actual inter-
actions between cultures and their environments. Godeman (2008)
terms this ‘Integration-oriented cooperation and boundary-crossing’.
For example, ethnographic and anthropological studies have shown
that current agriculture in mountainous Romania, as discussed in
Chapter 7, is strongly influenced by the cultural, political, and eco-
nomic history of the region, and ecologists have shown the impor-
tance of climate, topography, and geology. Less studied, however,
is how agricultural systems are formed by intertwined processes of
land use and vegetation, between societal needs and the landscapes’

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
3
8
2

Historical Ecology in Theory and Practice 283

provisioning capacity, through constant feedback between human


action and ecological response, and back to modified actions. Local
knowledge about such processes is extensive and adjustable; thus
‘traditional ecological knowledge’ serves to keep farmers prepared
and agriculture flexible enough to cope with dynamic culture-nature
interactions.
Research on such interaction processes often requires deep
insights and detailed analyses in both the natural and social sciences.
The research therefore needs interdisciplinary groups of research-
ers working closely together in a way that enables a lively exchange
among disciplines and beyond them, and an invitation to collabo-
rate with practitioners. In this book, we discuss several analytical
frameworks which can be used to work in historical ecology, such as
biocultural diversity, the careful borrowing of concepts, the careful
examination of baselines, and the attention to forms of time and to
traditional ecological knowledge. With the choice of a single frame-
work there follows focus and exclusion; certain aspects will be vis-
ible within one framework and not within another. Through close
collaboration and deep dialogue, we become aware of what the com-
bination of frameworks makes visible and to what we might still
be blind.
The exegesis research process will frequently require new types
of historical and ecological knowledge than what is usually utilised.
Some of this novel input can be obtained by new interdisciplinary
analysis, but, in many cases, such input can only be provided in
terms of non-academic knowledge, for example the experience and
skills of local farmers, livestock- and reindeer herders, and fisher-
men. Historical-ecological research therefore strongly benefits
from involving knowledgeable non-scholars, i.e., transdisciplinary
approaches, as discussed in Chapter 8.

9.4.1 New Sources of Knowledge


Historical sources often provide information about the state and use
of landscapes that can be interpreted in terms of ecological processes

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
4
8
2

284 Lennartsson et al.

and conditions, and thus contribute to our knowledge about the ecol-
ogy of species and habitats, both in the past and in the present. Such
interpretation requires careful examination of historical sources and
detailed ecological insights.
Another aspect of this type of knowledge is that if we under-
stand the causal links in how human activities have contributed to
forming ecosystems, we can also interpret current biodiversity as
a historical source. This would offer new ways of filling the land-
scape with historical information, making it an archive of biocul-
tural heritage that tells about land use and other activities in the
past (Rackham 1976; Barthel et al. 2013a, 2013b; Rotherham 2015).
Historical information can be derived from biodiversity at different
scales (Ljung et al. 2015):

• Traits and genetic material, such as varieties and breeds of domesticated


plants and animals, as well as adaptations to land use among non-
domesticated species, for example the evolution of early-flowering
ecotypes in hay meadows.
• Individual woody plants, i.e., trees and shrubs that are shaped by human
activities, either deliberate pruning, pollarding, etc. or as the result of
changed light conditions through human influence on the surrounding
forest stands (Figure 9.1).
• Populations of species that have historically been introduced or
established spontaneously in human-influenced habitats (cf. niche
construction in Chapter 6), and that persist long after the land use has
changed.
• Biotopes and vegetation types that are shaped by human activities; their
structure and species composition reflect earlier land use (Figure 9.2).
• Landscapes, in which contents of biotopes and elements reflect earlier
land use, for example the local land-use system and its needs for different
resource-provisioning ecosystems (Lennartsson et al. 2016).

The term biocultural diversity is used somewhat different con-


texts in the literature. It can be used to denote living biological carri-
ers of information about past human activities, as in the list provided
earlier, thus as a complement to other biological (e.g. pollen and mac-
rofossils), immaterial, and material sources of information frequently

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
5
8
2

Historical Ecology in Theory and Practice 285

Figure 9.1 Giant pollarded beech trees in a dense forest reveal the
land-use history as well as the former forest structure. This forest in
Botiza in the Romanian Carpathians has probably been a pasture with
scattered pollards, harvested for leaf fodder. A black-and-white version
of this figure will appear in some formats. For the colour version, please
refer to the plate section.
Photo: Tommy Lennartsson.

used by archaeologists. The term is also used to denote the actual


relationships between local communities and their environment, for
example in food production and rural development (Jonas, Bavikatte,
& Shrumm 2010; Argumedo & Swiderska 2014; Gavin et al. 2015).

9.4.2 Collaboration in Practice


In Chapter 8, several practical aspects of collaboration are dis-
cussed. There is a growing body of research in which ecologists have
applied some historical perspective to the analyses, or historians
have interpreted historical sources against an ecological background.
Usually, the ‘extra-disciplinary’ aspect is treated rather shallowly in

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
6
8
2

286 Lennartsson et al.

Figure 9.2 In the pre-industrial Swedish landscape, it was common


that midfield islets, used for grazing, were fenced together with the
arable fields. These islets – low hills – could not be grazed until after
harvest (August–September), and that grazing regime favoured plant
species that flower early and set seeds before grazing. One example is
the cow slip, Primula veris, whose presence indicates the historical
timing of grazing, long after management has changed. Hjälmö,
Stockholm archipelago, Sweden.
Photo: Anna Westin.

publications, and most articles are written in a style that is typical


for the research field of the publishing journal, either history or ecol-
ogy. Deeper interdisciplinary collaboration, as suggested previously,
will result in studies which require rather detailed descriptions of
the methods and sources used by both (or more) disciplines. Such
studies do not easily fit into conventional print outlets, and it may
therefore be difficult to publish in traditional ecological and histor-
ical journals. Even ecological and historical journals claiming to wel-
come interdisciplinary manuscripts often lack routines for offering
relevant interdisciplinary peer review or judgement, or flexibility to
include treatment of data and sources that are untypical for the jour-
nal (Ingerson 1994: 43–66). We believe that much historical-ecological

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
7
8
2

Historical Ecology in Theory and Practice 287

research is hampered by the lack of publishing possibilities, and that


there is a strong need for a specific journal of historical ecology, and
for more interdisciplinary journals in general.

9.5 Preserving Biodiversity and


Cultural Heritage
In the international conservation arena, there is an increasing aware-
ness of the need for holistic perspectives and knowledge about the rela-
tionships and intersects between nature and culture. IUCN declares
that human diversity through culture-based conservation and the con-
servation of species and ecosystems are ‘profoundly linked realities’
(2008: 49), and UNEP states that cultural diversity ‘can be affected by
the same drivers as biodiversity, and which has impacts on the diversity
of genes, other species, and ecosystems’ (2007). In order to apply these
general notions in practice, we need considerably deepened knowledge
about which of those links and drivers are most important for main-
taining the Earth’s capital of biodiversity and cultural heritage.

9.5.1 Historical-Ecological Approaches


in Conservation Planning
The importance of history is obvious in all conservation planning,
since conservation goals and activities cannot be designed unless we
know by what processes landscapes have been formed and changed,
as discussed in Chapter 5 (see also Rotherham 2015). In practice,
rather detailed information about historical-ecological processes and
conditions is often needed for connecting the status of species and
habitats with environmental conditions, and linking these condi-
tions to the history of land use and other anthropogenic or natural
processes. Such information is needed both for biodiversity conser-
vation and for conservation of the cultural heritage represented by
species, vegetation, and habitats in cultural (or ‘domesticated’, as dis-
cussed in Chapter 6) landscapes. An important task for applied his-
torical ecology is therefore to develop frameworks and methods for
analysing how historical human activities have changed ecosystems

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
8
2

288 Lennartsson et al.

and to translate the results of such analyses into practically useful


recommendations for landscape management.
From an applied conservation perspective, historical-ecological
analyses of this kind need to rest on three fundamental pillars of
knowledge: (1) which, if any, human activities have taken place in a
landscape; (2) how those activities have influenced species, habitats,
and landscapes; and (3) to what extent earlier human-environment
interactions still influence the present state of biodiversity. The first
two are derived by connecting human activities to ecological pro-
cesses in enough detail, and the third by connecting the past and the
present. The three pillars show that neither historical nor ecological
knowledge alone can inform conservation planning, but both need to
be analysed in the context of each other.
The pillars (1)–(3) indicate that knowledge about past human-
environment interactions is not equally important in every case of
conservation planning. Although human presence can be proven in
almost every corner of the world, not all of that presence has neces-
sarily influenced the natural environment significantly, and even sig-
nificant historical influence may have vanished to a degree that it
is no longer important for conservation goal-setting and planning.
Old-growth forests of conservation concern, for example, may carry
historical or archaeological traces of humans, but are now predom-
inantly characterised by natural ecological processes, that is, by the
absence of human activities.
It cannot be excluded, however, that even subtle traces of past
human activities represent anthropogenic ecological conditions that
are crucial for biodiversity. Without historical-ecological knowledge
about ecosystems, such ecological key factors risk being overlooked
by conservation planners (Rotherham 2007). Returning to old-growth
forests, in many regions, a large proportion of the species of conser-
vation concern are dependent on the few scattered habitat elements
that remain from historical forest grazing, such as veteran trees,
gaps, and patches of herb-rich vegetation (Franc et al. 2007). Such
elements also constitute an important cultural heritage of the forests

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
9
8
2

Historical Ecology in Theory and Practice 289

(Agnoletti et al. 2008). Historical-ecological analysis may indicate


that management activities are needed to favour such elements, or
even that forest grazing should be resumed, instead of minimising
human impact.
Conservation of threatened species aims at removing unde-
sirable environmental conditions and reintroducing desirable ones
(Caughley 1994). While the identification of threats can be done by
observing declining populations in their suboptimal environment,
it is often more difficult to identify the desirable conditions, since
they are no longer present to be observed in the field. By making an
ecological interpretation of historical information, for example about
land use, and an historical interpretation of the species’ ecology, such
‘lost’ key factors for population viability can be identified.

9.5.2 Dealing with the Multidimensional Past


‘The past’ is not a homogenous state, and we therefore need to con-
sider how the current ecological state may be the combined legacy
of several time periods. Any landscape carries traces of conditions
decades, centuries, and millennia ago, as well as those of yesterday
(Barak et al. 2016; see Chapter 2), and any reference point for con-
servation goal-setting (see Chapter 5) in changing landscapes will
thus consist of multiple time layers. In order to disentangle different
periods of human-environment interactions in a landscape, analy-
sis, including the pillars (1)–(3), can be performed. Such analyses are
interdisciplinary processes; archaeologists, historians, and long-time
residents of the area can provide understanding of past human activi-
ties in several dimensions. Ecologists can interpret the potential
impacts on ecosystems of those activities, and show which impacts
influence the present biodiversity. And the interdisciplinary dialogue
can provide insights in how humans and ecosystems have depended
on each other during different time periods. It is particularly impor-
tant to examine all periods of the past and not simply one or two
that might be attractive conservation goals, or have rich historical
information. For example, early travellers and scholars of American

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
0
9
2

290 Lennartsson et al.

environmental history neglected entirely or only fleetingly treated


significant and sophisticated Native American landscape modifica-
tion before the arrival of Europeans (Denevan 1992).
The multiple periods of time represented in landscapes usually
require simplification in practical conservation. Nature conservation
therefore often aims at solely ecological states of nature, which are
a mix of landscape elements from different time periods and would
almost invariably contain evidence from different periods of human
activity. Although such an approach may optimise the conditions for
current biodiversity in a reserve, it may aim at an unstable state, with
ecosystem elements that in the long run are not possible to combine.
Predictions of the long-term outcome of different conservation alter-
natives are facilitated by a historical-ecological awareness of how
subsistence systems and their land use have constituted historically
distinct ecological disturbance regimes that shaped the ecosystems
of different time periods. Choosing components (species, ecosystems)
from different time periods fragments the internal coherence of each
period, leaving the restoration work without context.

9.5.3 Identifying the Essential Components of the Past


A negative trend for biodiversity conservation status indicates that
the present environmental state is less favourable than an earlier,
‘before-decline state’, and biodiversity conservation therefore often
aims at restoring a certain past state of ecosystems (see Chapter 5).
Past conditions are, however, rarely possible (or desirable) to recon-
struct as a whole, but we need to identify those details of historical
conditions that are necessary for biodiversity (Plieninger, Höchtl, &
Spek 2006). Such details may be certain land-use activities (such as
mowing of hay meadows or grazing of pastures), or even certain com-
ponents of those activities (such as the intensity of grazing, and the
frequency of pollarding of trees; see Chapter 7). With proper know-
ledge of the necessary ecological functions of these land-use activities
of the past, we can then reintroduce them (e.g. based on traditional
ecological knowledge; see Chapter 4), or imitate them by using

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
1
9
2

Historical Ecology in Theory and Practice 291

modern methods, and if necessary modify the individual historical


components to fit changed environmental and societal conditions.
The fact that the present biodiversity may be dependent on eco-
system conditions that are no longer sufficient, if at all present, relates
to the concept of extinction debt, used in conservation biology (e.g.
Diamond 1972; Kuussaari et al. 2009). Some species can remain in
their habitats for a long period of time in spite of unfavourable envir-
onmental conditions, and can thus be considered to represent the
past. The extinction debt implies a potential: as long as species ‘from
the past’ are present, habitats can be restored. But it is also a problem
because conservation actions are needed in order to prevent further
species loss, also in habitats that are seemingly stable. When lost, it is
rare that species with low mobility come back spontaneously, even if
the environmental conditions have become favourable again.

9.6 Traditions and Traditional


Ecological Knowledge
As discussed in Chapters 4 and 7, humans’ use of natural resources
in different biomes on Earth has generated and is dependent on
extensive knowledge about ecosystems. This ‘traditional ecological
knowledge’ (TEK) is applied as land-use traditions, which are locally
adapted to both the environment and to culture and society. Today,
traditions and TEK are increasingly appreciated, both as cultural her-
itage and because they are needed when building a sustainable future.
Regarding the cultural heritage aspects of traditions and TEK,
traditions and customs related to land use have been recorded for
centuries, but, as Chapter 4 discusses, by involving also the cultural-
ecological aspect of the traditions, new dimensions of this cultural
heritage unfold. One characteristic feature of traditions is that they
constantly change, adapting to new environmental and societal con-
ditions. This further highlights the need to understand the traditions
both regarding their importance for land use and their cultural mean-
ing. We are convinced that new insights in historical environmental
conditions and human-environment relationships can be revealed

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
2
9

292 Lennartsson et al.

by interpreting traditions and TEK using a historical-ecological


framework.
Traditional ecological knowledge is needed in most types of
policy and planning that deal with the use of natural resources (e.g.
Berkes, Colding, & Folke 2000; Gomez-Baggethun, Ciorbera, & Reyes-
Garcia 2013). For restoring deteriorated environments and for learn-
ing from the past, we may need TEK no longer in use, but captured
in archival records. But for linking the past to the present, we need
to understand how TEK has changed along with changes in society
and the environment, and how it combines human-nature interac-
tions with culture. In such cases, we need TEK to be preserved in situ,
to allow it to change and adapt. Practices about how to use and relate
to nature may have their roots in the distant past of the studied time,
but may also be recently adapted to some societal change or technical
innovation (Mahoney 2000; Magnusson & Ottosson 2009). The term
traditional is in its essence vague as to which time that is concerning,
which makes many historians reluctant to use the term itself.
Memories of past practices and knowledge can be used as an
historical source, through interviews with local informants. As dis-
cussed in Chapter 5, memory is not the same as documentary evi-
dence of past times. Put differently, memories and documents entail
different source critical aspects. Memory is not objective recollection
of the past, but is largely formed by the present. The information
in documents is influenced by the purposes of the documents’ cre-
ation in a past time. As always when different sources give different
answers to a research question, the discordance should be taken as a
call for reflection on why the expected results did not turn up. Could
it be that the questions were not asked in an optimal way, or that the
method needs to be modified? In many cases, contradictory indica-
tions result in new insights and unexpected knowledge.

9.7 Sustainable Use of Natural Resources


In the past and today, shifting uses of natural resources strongly influ-
ence the dynamics of landscapes and ecosystems. From the latter half

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
3
9
2

Historical Ecology in Theory and Practice 293

of the twentieth century, many parts of the world have seen intensifi-
cation and expansion of agriculture and forestry in landscapes where
that production is profitable, and cessation of low-input agriculture
and diverse agroforestry in low-productivity landscapes. Although
much of this resource-use change can be assumed to be driven by
efforts to increase local sustainability, it has led to extensive loss
of values which are today seen as globally essential for sustainabil-
ity, such as biodiversity, cultural heritage, resilient ecosystems,
and global-scale ecosystem services, for example those regulating
greenhouse gases. This indicates that one important aspect of sus-
tainability from a historical-ecological perspective is that there are
considerable differences both between local sustainability and soci-
etal sustainability at broader societal scales, and between historical
and current components of sustainability.
Another essential historical-ecological aspect of sustainability
is the scale and degree of complexity of human-environment rela-
tionships. Before industrialisation, a large proportion of the inter-
actions occurred at local levels, where feedback mechanisms acted
efficiently since the culture’s effects on the environment were imme-
diate and visible. Local human-environment interactions included
not only land use and production in the landscape, but also the local
processing, selling, and consuming of comestibles. Today, many
more local products reach national and global markets, which has
removed much of the decision-making from the local ecosystem
users and thereby decoupled the consumers of food from the produc-
ers. It has also removed feedback mechanisms between production
and environmental status, and thus decoupled the products from the
provisioning ecosystems.
From a historical-ecological perspective, these changes can be
described as a conflict among scales. The globalisation of consump-
tion will most likely continue, which implies that resource use will
continue to be driven by broad-scale organisational structures far from
the natural resources that are central to production. Yet actual pro-
duction, the use of ecosystems, and the impacts on biodiversity will

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
4
9
2

294 Lennartsson et al.

remain local. These new means of resource use, including intensifi-


cation and homogenisation, are applied to landscapes and ecosystems
which are still ruled by the same ecological mechanisms as before.
The increasing distance between consumers and produc-
ers, between markets and ecosystems (Maffi & Woodley 2010: 4),
increases the risk of causing indirect, unintentional, and unexpected
environmental problems and decreases the possibility of adjusting
conditions of resource use fast enough (IUCN 2008). One important
task is to reconstruct some necessary feedback mechanisms between
users and natural resources. This requires that we trace links and feed-
back loops through complex markets and institutional organisations,
in order to identify the key links for the interdependence of humans
and nature in the globalised world. Another task is to find ways to
develop local traditional practices instead of abandoning them. This
requires identification of the most important components of human-
environment relationships, for example the key ecosystem services,
land-use practices, and institutions. These components need to be
developed in ways that both keep their role for sustainable resource
use and meet the needs for local change in a rapidly changing world.

References
Agnoletti, M., Anderson, S., Johann, E., Kulvik, M., Saratsi, E., Kushlin, A., Mayer,
P., Montiel, C., Parrotta, J., & Rotherham, I. D. (2008). The introduction of
cultural values in the sustainable management of European forests. Global
Environment, 2, 172–93.
Argumedo, A. & Swiderska, K. (2014). Biocultural Heritage Territories.
London: International Institute for Environment and Development.
Barak, R. S., Hipp, A. L., Cavender-Bares, J., Pearse, W. D., Hotchkiss, S. C., Lynch,
E. A., Callaway, J. C., Calcote, R., & Larkin, D. J. (2016). Taking the long
view: integrating recorder, archaeological, paleoecological, and evolution-
ary data into ecological restoration. International Journal of Plant Science,
177: 90–102.
Barthel, S., Crumley, C., & Svedin, U. (2013a). Bio-cultural refugia: safeguarding
diversity of practices for food security and biodiversity. Global Environmental
Change, 23(5), 1142–52. http://dx.doi.org/10.1016/j.gloenvcha.2013.05.001.

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
5
9
2

Historical Ecology in Theory and Practice 295

Barthel, S., Crumley, C., & Svedin, U. (2013b). Bio-cultural refugia: combating the
erosion of diversity in landscapes of food production. Ecology and Society,
18(4), 71. http://dx.doi.org/10.5751/ES-06207-180471.
Berkes, F., Colding, J., & Folke, C. (2000). Rediscovery of traditional ecological
knowledge as adaptive management. Ecological Applications, 10, 1251–62.
Caughley, G. (1994). Directions in conservation biology. Journal of Animal
Ecology, 63, 215–44.
Crumley, C. L. (1993). Analyzing historic ecotonal shifts. Ecological Applications,
3(3), 377–84.
Crumley, C. L. & Marquardt, W. H., eds. (1987). Regional Dynamics: Burgundian
Landscapes in Historical Perspective. San Diego, CA: Academic Press.
Denevan, W. M. (1992). The pristine myth: the landscape of the Americas in
1492. Annals of the Association of American Geographers, 82,(3), 369–85.
doi: 10.1111/j.1467–8306.1992.tb01965.x.
Diamond, J. M. (1972). Biogeographic kinetics: estimation of relaxation times
for avifaunas of southwestern Pacific islands. Proceedings of the National
Academy of Sciences, USA, 69, 3199–3203.
Egan, D. & Howell, E. A., eds. (2001). The Historical Ecology Handbook: A
Restorationist’s Guide to Reference Ecosystems. Washington, DC: Island
Press.
Ellen, R. (1982). Environment, Subsistence and System: The Ecology of Small-
Scale Social Formations. Cambridge: Cambridge University Press.
Eriksson, O., Bolmgren, K., Westin, A., & Lennartsson, T. (2015). Historic hay
cutting dates from Sweden 1873–1951 and their implications for conservation
management of species-rich meadows. Biological Conservation, 184, 100–7.
Franc, N., Götmark, F., Økland, B., Nordén, B., & Paltto, H. (2007). Factors and
scales potentially important for saproxylic beetles in temperate mixed oak
forest. Biological Conservation, 135, 86–98.
Gavin, M. C., McCarter, J., Mead, A., Berkes, F., Stepp, J. R., Peterson, D., & Tang,
R. (2015). Defining biocultural approaches to conservation. Trends in Ecology
and Evolution, 30, 140–5.
Godemann, J. (2008). Knowledge integration: a key challenge for transdisciplinary
cooperation. Environmental Education Research, 14, 625–41. doi: 10.1080/
13504620802469188.
Gomez-Baggethun, E., Ciorbera, E., & Reyes-Garcia, V. (2013). Traditional ecologi-
cal knowledge and global environmental change: research findings and policy
implications. Ecology and Society, 18, 72. doi: 10.5751/ES-06288-180472.
Gunn, J. & Crumley, C. L. (1991). Global energy balance and regional hydrology:
A Burgundian case study. Earth Surface Processes and Landforms, 16, 1–14.

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
6
9
2

296 Lennartsson et al.

Gunn, J., Crumley, C., Jones, E., & Young, B. K. (2004). A landscape analy-
sis of Western Europe during the early middle ages. In Charles L. Redman,
Steven R. James, Paul R. Fish, and J. Daniel Rogers, eds., The Archaeology of
Global Change: The Impact of Humans on their Environment. Washington:
Smithsonian Books, pp. 165–85.
Helldin, J.-O. & Lennartsson, T. (2007). Agricultural landscapes in East Europe
as reference areas for Swedish land management. In V. Surd & V. Zotic, eds.,
Rural Space and Local Development. Cluj-Napoca: Cluj University Press,
pp. 367–70.
Hoffman, S. M. & Oliver-Smith, A., eds. (2002). Catastrophe & Culture: The
Anthropology of Disaster. Santa Fe, NM: School of American Research Press.
http://dx.doi.org/10.5751/ES-06207-180471.
Ingerson, A. (1994). Tracking and testing the nature/culture dichotomy in practice.
In C. Crumley, ed., Historical Ecology: Cultural Knowledge and Changing
Landscape. Santa Fe, NM: School of American Research Press, pp. 43–66.
IUCN. (2008). Report of the Director General on the Work of the Union since
the IUCN World Conservation Congress, Bangkok, 2004. IUCN World
Conservation Congress, Barcelona, Spain 2008. http://cmsdata.iucn.org/
downloads/cgr_2008_8_dg_report.pdf.
Jonas, H., Bavikatte, K., & Shrumm, H. (2010). Community protocols and
access and benefit sharing. Asian Biotechnology and Development Review,
12, 49–76.
Jones, E. A. (2006). Surviving the Little Ice Age: Family Strategies in the Decade of
the Great Famine of 1693–1694 as Reconstructed through the Parish Registers
and Family Reconstitution. Chapel Hill: Dept. of Anthropology, University of
North Carolina.
Kellert, S. H. (2008). Chaos Theory and the Challenge of Learning across
Disciplines. Chicago: University of Chicago Press.
Kuussaari, M., Bommarco, R., Heikkinen, R. K., Helm, A., Krauss, J., Lindborg,
R., Öckinger, E., Pärtel, M., Pino, J., Roda, F., Stefanescu, C., Teder, T., Zobel,
M., & Steffan-Dewenter, I. (2009). Extinction debt: a challenge for biodiversity
conservation. Trend in Ecology and Evolution, 24, 564–71.
Le Roy Ladurie E., Daux V., & Luterbacher J. (2006). Le climat de Bourgogne et
d’ailleurs XIVe-XXe siècle. Histoire, économie & société, 3, 421–36.
Lennartsson, T., Westin, A., Iuga, A., Jones, E., Madry, S., Murray, S., & Gustavsson,
E. (2016). ‘The meadow is the mother of the field’: comparing transformations
in hay production in three European agroecosystems. Martor, 21, 103–26.
Ljung, T., Lennartsson, T., & Westin, A. (2015). Inventering av biologiskt kultur-
arv. Stockholm: Riksantikvarieämbetet.

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
7
9
2

Historical Ecology in Theory and Practice 297

Ljungqvist, F. C., Krusic, P. J. & Sundqvist, H. S., Zorita, E., Brattström, G., &
Frank, D. (2016). Northern Hemisphere hydroclimatic variability over the past
twelve centuries. Nature, 532, 94–8.
Madry, S., Jones, E. A., Murray, D. S., Tickner, A., & Misner, T. (2015). Water
and landscape dynamics in southern Burgundy: three and a half centuries of
water management in an agricultural landscape. Journal of Water History,
7(3), 301–35.
Maffi, L., & Woodley, E. (2010). Biocultural Diversity Conservation: A Global
Sourcebook. London and Washington, DC: Earthscan.
Magnusson, L. & Ottosson, L., eds. (2009). The Evolution of Path Dependence.
Cheltenham: Edward Elgar.
Mahoney, J. (2000). Path dependence in historical sociology. Theory and Society,
29, 507–48.
Newell, B., Crumley, C., Hassan, N., Lambin, E., Pahl-Wostl, C., Underdal, A., &
Wasson, R. (2005). A conceptual template for integrative human-environment
research. Global Environmental Change, 15, 299–307. doi:10.1016/
j.gloenvcha.2005.06.003.
Plieninger, T., Höchtl, F., & Spek, T. (2006). Traditional land-use and conservation
in European rural landscapes. Environmental Science and Policy, 9, 317–21.
Rackham O. (1976). Trees and Woodland in the British Landscape. Archaeology
in the field series. London, J. M. Dent & Sons Ltd.
Rotherham, I. D. (2007). The ecology and economics of medieval deer parks. In
I. D. Rotherham, ed., The History, Ecology and Archaeology of Medieval Parks
and Parklands. Sheffield: Wildtrack Publishing, pp. 86–102.
Rotherham, I. D. (2015). Bio-cultural heritage and biodiversity: emerging para-
digms in conservation and planning. Biodiversity Conservation, 24, 3405–29.
UNEP. (2007). Global Environment Outlook 4. Nairobi: UNEP.
United Nations. (2015). Sendai Framework for Disaster Risk Reduction 2015–2030.
A/RES/69/283 Annex II. www.unisdr.org/files/43291_sendaiframeworkfordrren
.pdf Accessed 24/02/17.

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:20, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.009
8
9
2

10 Taking Research into Action


in Historical Ecology
Carole L. Crumley

10.1 Introduction
We are living in a world where something we hold dear is swept away
every moment. Tumultuous news cycles report efforts to eradicate
entire human populations, bury their accomplishments with their
dead, erase them from their landscapes, and write them out of his-
tory. Less spectacularly and more stealthily, past worlds disappear
every day. A poem on a floppy disk goes into the trash, the last of a
key species dies, an ancient coastal fishing site slumps into high seas,
an excellent technique of water management meets the bulldozer.
Should we care? We already have more things to curate than
we can manage. Old things and ideas are, by definition, out of date.
But as human impact deepens, humans may themselves join other
casualties of our collective behaviour.
And yet our species has a long creative history: it is more than
seventy thousand years since we became ‘human’ in mind and body.
Surely many earlier ideas could still be useful and even transforma-
tive, because they are the fruit of the same recurrent struggles with
materials, weather, and forms of society that are present today. Fresh
water, snug buildings, and productive fields were important long
before they became bottled water, cozy ski lodges, and ecosystem
services.
For millennia we have thrived by observing and taking action,
leaving behind a record of the directions we have taken. For human-
ity, the world has been a laboratory in which our species speculated
and experimented. Not all those directions were successful, but over
time we have learned new ways to address changed circumstances
and to evaluate our efforts. It is the passage of time that demonstrates

298

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
9
2

Taking Research into Action in Historical Ecology 299

what is variously called durability, robustness, or resilience and,


especially for the evaluation of practices, time is the only test.
If humanity is able to learn quickly, our species could continue
to flourish and we may still reach a new shore. But humanity may
already be aboard a planetary Titanic. If we must fill a few lifeboats
with messages for the future, what will they be? What forces – ideas,
conditions, inventions – drive change? What are quintessential
human needs beyond the physical? What redeeming arguments can
we leave, perhaps for no one, that showcase the pacific side of a spe-
cies with the conscious ability to destroy life on the planet? How,
both practically and ethically speaking, can we set out into space in
the search for other planets to ruin?
Researchers drawn to historical ecology aren’t ready to dis-
card a legacy which could help us shape a more desirable future in
a changed world. This legacy is not all about things. It includes an
intangible store of knowledge: how societies are and have been organ-
ised, how important information is transmitted from one generation
to another, and how to get along with each other individually and in
groups. This is the very fabric of society.
The social sciences and the humanities are storehouses of
information about how we can construct durable and meaningful
societies. In addition to practical knowledge, societies give us confi-
dence, emotional comfort, and a sense of belonging. Environmental
sciences, traditional knowledge, and other ways of knowing about
our surroundings offer understanding of a critical bond with the
natural world that supports us. In the riotous diversity of the living
and inanimate world we find both inspiration and insurance for the
future.
So the first thing to bring with us into the future should be a
guidebook for the territory that has already been charted. The flex-
ible toolbox of historical ecology enables us to learn how environ-
ments, practices, and policies interact over time. This book’s final
chapter will explain the importance of the past to sceptics, examine

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
0
3

300 Crumley

means by which the past can help meet enormous global challenges,
and offer ideas that are already available for charting a liveable and
satisfying future for our species and a healing prescription for Earth.

10.2 Critical Approaches and Values:


Historical Ecology’s Unique Perspectives
What aspects of historical ecology are especially valuable for the
future? Among the most important is helping to heal the rupture
between objectified ‘Nature’ and peoples’ lives; we must demon-
strate that the natural world which supports us still matters.
A growing river of knowledge carried our most distant human
ancestors into their future and our present. How was its value lost to
us, and when? Among influential movements in Western intellectual
history, two have much to say to the current quandary.
Ideas about the purpose of the world’s bounty were writ-
ten in the first chapter of the Hebrew Bible and the Christian Old
Testament, half a millennium before Christ, to guide the mobile and
restless populations of a semi-arid global crossroads. Fifteen hundred
years later they were used to justify global-scale colonial domination
over others’ lands and wealth.
The Western Enlightenment enshrines reason as the primary
source of authority, whence flow contemporary ideals of tolerance,
the separation of church and state, progress, and constitutional
government; reason is the cornerstone of Western science. The
Enlightenment ushered in scientific authority that brought longer
and healthier lives, the awesome inventions of flight, communica-
tions, and much more; it has also enabled environmental destruc-
tion that can endanger humanity at an unprecedented scale (e.g. the
atomic bomb, fracking, and other technologies).
These ideas, uncomfortable neighbours for centuries, nonethe-
less led both scholars and the general public to abjure not simply
traditional knowledge but almost any value, other than romantic,
of an unlightened past. Progress and new technologies can do nicely
without the past. We imagine technological salvation (space travel,

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
1
0
3

Taking Research into Action in Historical Ecology 301

clean energy) from consequences of the reckless use of chemicals,


bioengineering, nuclear materials, and fossil fuels. The ‘technology
solution’ now, confusingly, both gives us new forms of energy and
underwrites contemporary global agendas that threaten the planet
and impoverish humanity.
But humans are social animals and still need something to
believe in and communities to join. At all scales, the environmental
sciences investigate the web of life and the global ecosystem while
the social sciences and humanities explore human cognition and val-
ues; both seek to understand aspects of the extraordinary phenom-
enon that is Life. This book argues that it is possible to meld these
fields of study so that all available information – social, political,
environmental, emotional – is necessary for wise decision-making.
The future must be guided by both material and immaterial principles
that acknowledge and mediate human emotional needs, strengthen
the social fabric of communities everywhere, and give urgent atten-
tion to the environmental and other conditions that threaten life on
Earth.
Historical ecology values the knowledge held by communi-
ties, often termed traditional or local. As Chapters 4 and 8 in this
volume make clear, such communities have much to offer, along
with contemporary intentional communities of practice – River
Keepers, permaculturists and community gardeners, the Slow Food
movement – that address environmental issues and strengthen the
social fabric.
In placing fields of study and communities of practice together,
an important area of agreement is revealed. Place-based practice,
with its inherent local and regional focus, anchors historical ecol-
ogy and completes connections to the work of Earth system scien-
tists. Still needed, however, has been a flexible bottom-up approach
that connects local and regional environmental research with past
and present human activity and with the Earth system. Historical
ecology’s framework makes these new collaborations possible, link-
ing the study of broad planetary processes to knowledge of humans’

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
2
0
3

302 Crumley

interaction with dynamic landscapes. These new links are already


helpful in engendering scholarly thinking at multiple, connected
scales from local to global (e.g. Chapter 3).
The failure of the international governmental policy establish-
ment to make definitive progress in slowing the drivers of climatic
change has necessitated this grass-roots response. Partnerships and
networks that build consensus and promote collective action are
forming everywhere: cities and their mayors, groups that foster social
and environmental justice, co-housing communities, town councils
challenged to lower the cost of electricity or mitigate the effects of
climate change.
For these diverse networks to succeed, values (e.g. equity, trans-
parency) must be clearly articulated and the value of many kinds of
knowledge must be appreciated. Communities of both scholarship
(diverse disciplines, methods, theories) and practice (e.g. local/indig-
enous communities, farmers, hunters, managers, NGOs, town coun-
cils) must be included.
These alliances are crucial inasmuch as the past and the pre-
sent are valuable in the lives of individuals and constitute the fabric
of communities’ futures. Safeguarding biocultural heritage and pass-
ing on lessons and insights are quintessentially human activities and
must be in our waiting lifeboats.

10.3 Explaining the Past to the Present:


Solid Arguments
It is easy to think of the past as being, well, past. The pace of con-
temporary life is such that the over-and-finished easily loses out to
the about-to-happen. But several aspects of past events are of great
importance. There is the opportunity to find out how things turned
out: to study an historic crisis is to learn that the outcome was any-
thing but predetermined.
The investigation of quandaries, turning points, and plights
offers the possibility of evaluating, from our present position, how
such times were managed. While past events may not exactly match

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
3
0

Taking Research into Action in Historical Ecology 303

current predicaments, patterns and processes that are likely to recur


can be identified; they offer a laboratory for the meta-analysis of the
conditions which resulted in the dilemma.
In systems, cascade effects occur with the failure of a key ele-
ment (e.g. in ecology, a key species) or a crucial resource (water). The
possibility of such catastrophic failures increases in hyper-connected
systems, such as are found today; they have old and new vulnerabili-
ties (vulnerable power sources, contagious disease, adverse weather)
that risk managers must analyse and try to avert. This proactive risk
response approach is utilised in the management of particularly dan-
gerous systems such as nuclear power plants. It is possible to use
similar methods to measure risk and vulnerability in the past. The
past provides convincing evidence for precisely what constitutes
durability under certain conditions, by identifying what essential ele-
ments endured or failed, under what conditions, and in what interac-
tion with one another.
Historical ecology offers a wide variety of resources for better
contemporary management and increased well-being in the future.
These can be applied at all spatial scales, from that of the planet
to regions, bioclimatic zones, landscapes, and specific places. Next
are discussed the major areas in which historical ecology can help
safeguard natural resources; environmental and cultural heritage;
aesthetic and spiritual values; the preservation of archives; and the
stimulation of innovation.

Managing Natural Resources. Historically and ecologically informed,


pragmatic approaches to land use, land management, biodiversity,
and restoration undergird sustainable management strategies and
allow measurement of sustainability over time. Perz and Almeyda
(2010) provide an example in forest management. A considerable por-
tion of the world’s land is covered by secondary forests, from the
tropics to the polar bioregions. Perz and Almeyda examine the ways
secondary forests, reforestation, and forest recovery are linked. This
requires integrative theoretical approaches that are open to multiple

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
4
0
3

304 Crumley

explanations and more adaptable to specific contexts. The authors’


integrative approach for land use/land cover change draws on his-
torical, landscape, and political ecologies, and on several aspects of
complex systems theory. By offering a common framework in which
various explanations can be arranged and evaluated and that incor-
porates temporal variation, systematic comparisons can be made
while retaining place-based ecological and political particularity. In
this volume, the authors employ historical ecology to understand
the utility of historic strategies to improve contemporary grasslands
management (Chapters 4, 6, and 7).

Biocultural Heritage. The European Union–funded project


Mediterranean Mountainous Landscapes in Europe (www.memola
project.eu/) analyses the environmental impact of historical agricul-
tural land uses in order to generate conservation strategies. Working
in Spain, Albania, Italy, and Sicily, the project partners with local
residents apply time-tested strategies, such as the restoration of
ancient varieties of fruit trees to abandoned land, and the restor-
ation of the region’s medieval water management infrastructure,
once again in use today. By implementing and improving strategies
for these regions’ conservation and protection, the study re-values
cultural assets, enhancing both environmental sustainability and
local identity. The intangible cultural heritage of their communities
is introduced to young people by their elders, maintaining regional
traditions through shared knowledge and complementing the team’s
scholarly approach. Together they promote both local economic
development and social awareness (Chavarría Arnau & Reynolds
2015; Memola Project 2014). In this volume, Chapters 4 and 7 offer
similar examples.

Heritage and Environment: Their goals are entangled. The U.S.


National Parks System (NPS) has used historical ecology to guide
the interpretation and management of its sites and regions since
the 1990s. These often-contested properties benefit from historical

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
5
0
3

Taking Research into Action in Historical Ecology 305

ecology’s theoretical and ethical inclusivity; at stake are the cultural


identities of many traditional communities (e.g. Native American
homelands, early settler towns) and the needs and goals of a diverse
general public (NPS sites host a varied clientele: battle re-enacters,
hikers and climbers, river rafters, families with small children and
pets, visitors to the United States, researchers).
Visitors may be attracted to different characteristics of a site,
such as its natural beauty, its historical importance, or its cultural
value; thus in a holistic and pragmatic approach to site management,
it is necessary to adopt the premise that any site’s meaningful charac-
teristics are not either Nature or Culture but both, deeply entwined.
The bio- and geophysical world, and what our many minds
and hands make and have made of it, is a state of affairs that can
be thought of as entangled, networked, a web, a matrix, a complex
system, etc. Historical ecology offers a framework for practical sense-
making but declines to offer a single version of history: individuals
and groups see the world around them differently and constantly
rework the past to fit a shifting present. Different as our pasts may
be, we are all part of a tightly connected system; we must now refash-
ion old lessons about how to care for the planet that sustains us.
So why should collective efforts to cherish our world (such as
UNESCO World Heritage Sites, the Nature Conservancy, the World
Wildlife Fund) be fighting separate battles to preserve a single aspect
(wildlife, or historic buildings, or ancient forests), leading to fierce
competition for funding and too often the banishment of indigenous
populations? Humans have been telling and learning stories about
creatures, events, and places – all mixed together – for millennia;
narratives that link the listener or observer to the complex histo-
ries of places and other living things end up teaching all aspects of
conservation.
As U.S. National Parks Service Director Jon Jarvis told the U.S.
Senate in 2009, ‘One of the most precious values of the national parks
is their ability to teach us about ourselves and how we relate to the
natural world. This important role may prove invaluable in the near

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
6
0
3

306 Crumley

future as we strive to understand and adapt to a changing climate’


(quoted in Rockman et al. 2016).

Aesthetic and Spiritual Values. Robin Wall Kimmerer writes:


‘Traditional ecological knowledge . . . exists all over the world, inde-
pendent of ethnicity. It is born of long intimacy and attentiveness to
a homeland and can arise wherever people are materially and spiritu-
ally integrated with their landscape’ (2002). In traditional commu-
nities, the relationships among aesthetics, spirituality, and intimate
knowledge of the environment are manifold, and history is woven
into the fabric of daily life. One of the most striking examples of
this entanglement comes from anthropologist and linguist Keith
Basso’s work (1996; Feld & Basso 1996) with members of the White
Mountain Apache Tribe, whose Arizona landscape is filled with out-
crops, passes, and vistas that serve as mnemonics for past events that
reflect the group’s core values. Embedded in the traditional environ-
mental knowledge of their landscape are the respect, reciprocity, and
responsibility that are central to environmental ethics. ‘The concept
of dwelling assigns importance to the forms of consciousness with
which individuals perceive and apprehend geographical space. More
precisely, dwelling is said to consist in the multiple “lived relation-
ships” that people maintain with places, for it is solely by virtue of
these relationships that space acquires meaning’ (Basso in Feld &
Basso 1996: 55). Communities and individuals everywhere can find
material and spiritual benefits in places that encourage curiosity
and contemplation, care and responsibility. The White Mountain
Apaches have built a framework to safeguard their values and ensure
their communities’ future (http://ihopenet.org, search ‘Fort Apache’).
In this volume, Chapters 4 to 7 offer examples of these shared, deeply
layered relationships with physical surroundings.

Archives. Climate change is rapidly accelerating the loss of infor-


mation about the former state of Earth’s environments and about
the ways people managed them. Rising winds, ocean levels, and soil

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
7
0
3

Taking Research into Action in Historical Ecology 307

temperatures are impacting both human heritage and archives of


environmental data; archaeological sites and environmental stores,
such as peatlands and packrat middens that have endured for centu-
ries and millennia, are threatened around the world. Once exposed,
iconic environmental, historic, and prehistoric properties are rapidly
destroyed (Harvey & Perry 2015; Rockman et al. 2016). These sites
are resources for measuring long-term variability in key economic
species (e.g. cod, salmon, insects, crops) and past adaptation to cli-
mate change; they are being destroyed just as their full potential for
global change science is being realised.
Archaeological sites with good conditions of organic preserva-
tion support palaeoenvironmental reconstruction: their potential is
similar to stratified, datable proxy records recovered from ice sheets,
bogs, lakes, and oceanic sediments. This information can allow cor-
rection of current climate models and refine them to the regional
scale, enhancing the ability to anticipate future climates and to
respond to change (Chapter 3). They also form the basis for recon-
structing historic and prehistoric resource management. An example
of current efforts to save these precious resources is the Distributed
Long Term Observing Network of the Past (http://ihopenet.org
search ‘DONOP’). Thus in addition to piecing together the human
past, archaeological data can also be used to address questions that
fall primarily into the realms of conservation biology, oceanography,
ecology, and climatology (Sandweiss & Kelley 2012).

Innovation. Innovation is defined as a new idea, or a more effective


device or process. Countless innovations have influenced human his-
tory and many more are needed to steer a new course for human-
kind. Old ideas stimulate new ones: retrofits to water resources,
agriculture, architecture, and earth or stone-made structures such
as terraces (collectively termed landesque capital, see Håkansson &
Widgren 2014) yield new resources and new solutions. Place-based
management solutions – which begin with an evaluation of terrain,
exposure, vegetation, drainage, and other enduring features of the

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
8
0
3

308 Crumley

local environment and their historic management – can be sturdy,


inexpensive, relatively easy to maintain, and applicable to contem-
porary problems.
Thus the past is full of inventions, and not all of them have
been surpassed by contemporary ways of doing things. Even failures
are useful: an idea that was visionary but ultimately failed a few
years or centuries ago could – with today’s materials, techniques, and
perspectives – enjoy success. Many ideas from the past are character-
ised by great attention to their context, such as considering season-
ality when choosing materials for buildings and their placement, or
finding clever ways to irrigate a terraced landscape. Swiss mountain
communities have managed an elaborate terracing system for cen-
turies, with the aid of transparent rights and responsibilities devel-
oped collectively among families (Netting 1981). Water management
is particularly interesting for re-envisioning, in that earlier manage-
ment schemes were built with local materials and labour and good
knowledge of terrain, soil, and geology (Håkansson & Widgren 2014).
Recently archaeologists, in collaboration with hydrologists and engi-
neers, studied the evolution of the Maya civilisation’s water man-
agement system (400 BC–AD 900). The Maya managed Yucatan’s
semi-arid environment and difficult topography with techniques of
water collection and storage that have much to offer today’s cities in
similar environments (Chase & Scarborough 2014; Scarborough et al.
2012; Scarborough, Chase, & Chase 2012).
It’s not just materiality: old management strategies are often
(still) valuable today. Megan Hicks and her colleagues (Brewington
2015 et al.; Hicks 2016 et al.) have studied remains of thousand-year-
old duck egg harvests from archaeological sites near Lake Myvatn in
northern Iceland; they have also interviewed present-day Icelanders
and searched medieval Icelandic literature. Their conclusion is that
historic strategies for both utilising and maintaining the duck popu-
lation have remained the same up to the present. This confirms both
the sustainability of the practice and the importance of place-based
knowledge which is shared and passed on to the next generation.

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
9
0
3

Taking Research into Action in Historical Ecology 309

10.4 From Research to Action


How to transform historical ecology and similar efforts into collec-
tive action on behalf of fundamental human values and a liveable
future? Community-based initiatives are popular everywhere, but are
they being heard by decision-makers? How can scholars, accustomed
to speaking and writing to one another, join these efforts and facili-
tate the transmission of important information to policy-makers?
Most researchers are not prepared to meet these challenges.
Researchers in the bench sciences rarely have such experience. The
situation is somewhat better in the social sciences and applied pro-
fessions (e.g. anthropology, sociology, public health, law), where
is it routine to meet with residents of a place or clients for a ser-
vice. Nonetheless, relatively few academics have the skills, time, or
opportunities to engage elected representatives, international NGOs,
or corporate boardrooms.
Established scholarly organisations can foster the work of
researchers and practitioners. For example, at the international level,
a 1972 UN conference on the environment stimulated the global
change community1 to launch a grand effort to unite scholars across
environmental science disciplines. This undertaking revolutionised
Earth system science and organised a worldwide network of research-
ers who together trained the next generations of scholars.
Today another milestone is in sight. The International Council
for Science (ICSU) is a non-governmental organisation of national
scientific bodies and international scientific unions in 142 coun-
tries across the globe. Founded in 1931, it is one of the oldest non-
governmental organisations in the world. In the social sciences,
its counterpart is the International Social Science Council (ISSC).
Founded in 1952, it is also an independent and international NGO

1
The global change community in 1972 was comprised of the World Climate
Research Programme [1979–present], the International Geosphere-Biosphere
Programme [1987–2015], and the International Human Dimensions Programme
[1990–2014].

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
0
1
3

310 Crumley

and represents the social, economic, and behavioural sciences. In


October 2016 these groups voted to merge, vowing to respond to
global challenges by fostering cooperation among scholars and stake-
holders everywhere.
Recognising that contemporary challenges require a sweeping
redesign of the global scholarly community and the way it responds
to critical issues, ICSU had already joined with key funding agen-
cies to form the Belmont Forum (2009), which promotes engage-
ment between research funding agencies and the academic research
community through co-design, co-alignment, and co-funding
of major research programmes. The coalition was joined by the
United Nations Educational, Scientific, and Cultural Organization
(UNESCO [1945]).
Towards these ends, Future Earth, a ten-year international
research initiative, has been launched to develop the knowledge to
respond effectively to global environmental change and to support
global sustainability. Taken together, these international scholarly
communities seek to reconfigure global-scale institutions to meet
stark challenges at every geographic scale.
How might practitioners of historical ecology contribute their
skills amid this tangle of acronyms and big organisations? Many
researchers have already committed their time and energy to the
international effort, but it is clearly not for everyone and not enough.
A more effective use of historical ecology may be to introduce other
scales of analysis. The old global programmes were just that: global.
Their work reflects the planet in the aggregate, but it is at the ground
level that solutions to global issues emerge.
Eventually, sustainable management must be moulded to spe-
cific places; nothing replaces the careful study of how regional and
local conditions have arrived at their present state. Historical ecol-
ogy offers two important features: an already-in-place platform for
close collaboration across disciplines, and a commitment to on-the-
ground work with communities and coalitions that are fundamental
to human organisation.

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
1
3

Taking Research into Action in Historical Ecology 311

With a long-term, human-scale, and place-based perspective,


the Integrated History and Future of People on Earth (IHOPE; http://
ihopenet.org/) is a network of researchers and research projects that
integrates scholars with stakeholders. IHOPE’s research design com-
bines historical ecology, the environmental humanities, and future
studies (which includes complex adaptive systems, modelling, and
scenario planning). The IHOPE approach can link global to local
scales, facilitate pragmatic, sustainable, and ethical management
strategies, and ‘grow’ regional expertise to manage the future.
A robust example is the rapidly growing Circumpolar Networks
coalition (http://ihopenet.org/circumpolarnetworks/) which combines
the work of the Nordic Network for Interdisciplinary Environmental
Studies (www.nordforsk.org/en, search ‘NIES’), the North Atlantic
Biocultural Organization (www.nabohome.org/, search ‘NABO’), and
the Global Human Ecodynamics Alliance (GHEA; www.gheahome
.org/). The coalition has hundreds of members who are resident citi-
zen scientists and/or scholars of the circumpolar region; together they
identify appropriate and place-based solutions that reduce risk and
enhance sustainability.
These collaborations are predicated on the power of independ-
ent, place-based, grass-roots, big-and-small endeavours that share
values and tools for study. There are many more examples from
throughout the world; historical ecologists can join existing collabo-
rations or start their own, pooling resources, contacts, and ideas.
This emergent, collaborative, transdisciplinary, and trans-
temporal research environment can address critical issues facing
humanity. Among them are the role of biodiversity in food security;
the need to rethink the sustainability of contemporary engineering,
architecture, and cities; the future of regions; and the role of history
and memory in a changed world. Much to do, room for everyone.

References
Basso, K. (1996). Wisdom Sits in Places: Landscape and Language among the
Western Apache. Albuquerque: University of New Mexico Press, pp. 13–52.

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
2
1
3

312 Crumley

Brewington, S., Hicks, M., Edwald, Á., Einarsson, Á., Anamthawat-Jónsson, K.,
Cook, G., Ascough, P., Saáyle, K. A., Arge, S. V., Church, M., Bond, J., Dockrill,
S., Friðriksson, A., Hambrecht, G., Juliusson, A. D., Hreinsson, V., Hartman,
S., Smiarowski, K., Harrison, R., & McGovern, T. H. (2015). Islands of change
vs. islands of disaster: managing pigs and birds in the Anthropocene of the
North Atlantic. The Holocene, 25(10): 1676–84.
Chase, A. F. & Scarborough, V. F. eds. (2014). The Resilience and Vulnerability
of Ancient Landscapes: Transforming Maya Archaeology through IHOPE.
Archaeological Papers of the American Anthropological Association, vol.
24. Wiley.
Chavarría Arnau, A. & Reynolds, A. eds. (2015). Detecting and Understanding
Historic Landscapes. Mantova: SAP Società Archeologica SRL.
Feld, S. & Basso, K. eds. (1996). Senses of Place. Santa Fe, NM: School of American
Research Press.
Håkansson, T. & Widgren, M. eds. (2014). Landesque Capital: The Historical
Ecology of Enduring Landscape Modifications. Walnut Creek, CA: Left Coast
Press.
Harvey, D. & Perry, J. eds. (2015). The Future of Heritage as Climates
Change: Loss, Adaptation and Creativity. Key Issues in Cultural Heritage.
London, New York: Routledge.
Hicks, M., Einarsson, Á., Anamthawat-Jónsson, K., Edwald, Á., Friðriksson, A.,
Þór Þórsson, Æ., & McGovern, T. H. (2016). Community and Conservation:
Documenting Millennial Scale Sustainable Resource Use at Lake Mývatn,
Iceland. In C. Isendahl & D. Stump, eds., Handbook of Historical Ecology and
Applied Archaeology. Oxford: Oxford University Press.
Kimmerer, R. W. (2002). Weaving traditional ecological knowledge into biologi-
cal education: a call to action. Oxford Journals: Science & Mathematics,
BioScience, 52(5), 432–8.
Memola Project (2014). Mediterranean mountainous landscapes: an histori-
cal approach to cultural heritage based on traditional agrosystems. Agua y
Territorio, 4, 146–8. Universidad de Jaén, Jaén, Spain.
Netting, R. M. (1981). Balancing on an Alp: Ecological Change and Continuity
in a Swiss Mountain Community. Cambridge: Cambridge University Press.
Perz, S. G. & Almeyda, A. M. (2010). A Tri-partite Framework of Forest Dynamics:
Hierarchy, Panarchy, and Heterarchy in the Study of Secondary Growth. In H.
Nagendra & J. Southworth, eds., Reforesting Landscapes: Linking Pattern and
Process. Netherlands: Springer, pp. 59–84.
Rockman, M., Morgan, M., Ziaja, S., Hambrecht, G., & Meadow, A. (2016).
Cultural Resources Climate Change Strategy. Washington, DC: Cultural

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
3
1

Taking Research into Action in Historical Ecology 313

Resources, Partnerships and Science and Climate Change Response Program,


National Park Service.
Sandweiss, D. & Kelley, A. R. (2012). The archaeological record as a paleocli-
matic and paleoenvironmental archive. Annual Review of Anthropology, 21,
371–409.
Scarborough, V. L., Chase, A. F., & Chase, D. Z. (2012). Low-density urbanism,
sustainability, and IHOPE-Maya: can the past provide more than history?
Urbanization and Global Environmental Change Viewpoints, 8, 20–4.
Scarborough, V. L., Dunning, N. P., Tankersley, K. B., Carr, C., Weaver, E., Grizioso,
L., Lane, B., Jones, J. G., Buttles, P., Valdez F., & Lentz, D. L. (2012). Water and
sustainable land use at the ancient tropical city of Tikal, Guatemala. PNAS,
109(31), 12408–13.

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
4
1
3

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 23 Feb 2018 at 16:54:23, subject to the
Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780.010
5
1
3

Index

Action, taking research into, 8, 309–11 work exchanges and, 94–95


Adaptation yeomen in colonial America (See Yeomen
biocultural diversity and, 189, 226 in colonial America)
climate change, adaptation Akkadian Empire, climate change and, 56
strategies, 61–63 Albania, biocultural heritage in, 304
grasslands and, 165 Alliances with communities of practice
revisiting, 153–55 (CoP), 302
summer farming in Sweden and, 226 Almeyda, A.M., 303–04
yeomen in colonial America and, 226 Anasazi people, climate change and, 56
Aesthetic values, 306 Annalistes, 17–19
Africa Anthropocene Era, 1, 117
domesticated landscapes in, 156–62 Anthropological Contributions to Historical
glades in, 157–59, 161–62 Ecology: 50 Questions, Infinite
livestock pens in, 159–60, 167–69 Prospects (Armstrong et al.), 9
umbrella thorn in, 160–61 Applied biocultural diversity, 228–30
African Dark Earths (AfDEs), 157 Archaeology
Agar, M., 23 archives and, 307
Agriculture communication in, 248
Africa, livestock pens in, 159–60 dependence in, 149–50
climate change and, 60–61 entanglement in, 149–50
common resources and, 92–94 New Archaeology, 15, 22–23
in England, 228 radiometric dating in, 15
feasts and, 96–97 stratigraphy in, 14
in France, 227–28 time in, 14–15
grassland-based agriculture in Romania uniformitarianism in, 14–15
(See Grassland-based agriculture in Archives, 306–07
Romania) Arctic amplification, 46
mowing and, 96 Ariadne (Greek mythology), 2
reindeer husbandry in Sweden (See Aristotle, 13, 16, 207
Reindeer husbandry in Sweden) Arizona, White Mountain Apache Tribe
risk management in, 227–28 in, 306
in Romania, 102–03 (See also Grassland- Armed conflict, climate change and,
based agriculture in Romania) 57–58, 65–66
seasonality and timing and, 95–97 Atran, S., 134–35
summer farming in Sweden (See Summer Authorship, collaborative research
farming in Sweden) and, 252–55
in Sweden, 101–02 (See also Reindeer
husbandry in Sweden; Summer farming Backcasting, 279
in Sweden) ‘Bad room’, 249–50
traditional ecological knowledge (TEK) in Baffin Island (Canada), pristine
rural communities and, 85, 91–92 environments in, 118
weather prediction and, 97–100 Bailey, G., 32

315

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:54:22, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
6
1
3

316 Index

‘Balance of nature’, 150 new sources of knowledge and, 284


Barad, Karen, 167 pastures and, 224–25
Baselines, 6–7, 112–14, 134–35 preservation of, 287–91
biocultural awareness and, 116–20 in reindeer husbandry in Sweden (See
biological diversity and, 115–16 Reindeer husbandry in Sweden)
in conservation, 114–16 similarities between biological and
defined, 114 cultural diversity, 189–91
‘extinction of experience’ and, 134–35 spatial and temporal complexity
historical awareness and, 6–7, 116–20 and, 224–25
in marine sciences, 114–15 in summer farming in Sweden (See
pristine environments and, 116–20 Summer farming in Sweden)
setting of, 120–22 Biocultural heritage, 304
Shifting Baseline Syndrome (SBS) (See Biological diversity. 7, 184. See also
Shifting Baseline Syndrome (SBS)) Biocultural diversity
as socially constructed, 112–13, 113n. 1 among yeomen in colonial America,
Basic questions concerning historical 199–206 (See also Yeomen in colonial
ecology, 4–5 America)
Basso, Keith, 306 baselines and, 115–16
Bateson, Gregory, 21 biotopes and vegetation and, 284
Bauzot (France), ponds in, 260–61 buffering and, 185–86
Belmont Forum, 310 causes of, 184–85
Belvedresi, R.E., 129 competition for resources and, 185, 190
Biblical perspective on ecology, 300 differences with cultural diversity, 188–89
Biocultural diversity, 7, 187–88 differing perspectives on, 191–92
adaptation and, 189, 226 ecological and evolutionary age and, 184
among yeomen in colonial America, effects of, 185–86
199–206 (See also Yeomen in colonial food availability, water availability, and
America) temperature and, 184
applied biocultural diversity, 228–30 gradients of, 184
concept of, 187–88 in grassland-based agriculture in Romania
conservation and, 304 (See Grassland-based agriculture in
correlations versus mechanisms Romania)
in, 222–23 human-nature relationships,
defined, 187–88 understanding mechanisms of, 192–93
differences between cultural and interaction with cultural diversity, 189–91
biological diversity, 188–89 landscapes and, 284
at different levels, 223–24 new sources of knowledge and, 284
differing perspectives on, 191–92 populations and, 284
drivers of, 224–26 predation and, 185
entanglement and, 304–06 predictability, stability, and continuity
extensions of, 221–22 and, 184, 185–86
in grassland-based agriculture in Romania preservation of, 287–91
(See Grassland-based agriculture in in reindeer husbandry in Sweden (See
Romania) Reindeer husbandry in Sweden)
historical ecology and, 226–28 resilience and, 185, 189
human-nature relationships, in Romania, 105–06
understanding mechanisms of, 192–93 seasonality and, 184–85, 190
implications of, 221–22 similarities with cultural
interaction between biological and diversity, 189–91
cultural diversity, 189–91 spatial heterogeneity and, 185, 190

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:54:22, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
7
1
3

Index 317

in summer farming in Sweden (See in Central Asia, 56


Summer farming in Sweden) in China, 46–47, 51, 56, 57–58
traditional ecological knowledge (TEK) in current status of, 54–55
rural communities and, 105–06 cyclical variability, 45–46
traits and genetic material and, 284 Dark Age Cooling Period, 47, 56
woody plants, 284 dendroclimatology and, 48
Biology de Vries cycle, 45–46
adaptation, revisiting, 153–55 documentary evidence, 48
niche construction, revisiting, 151–53 in England, 63
time in, 16 environmental determinism and, 57
Biotopes, biological diversity and, 284 failure to respond to, 302
Bloch, Marc, 17, 227–28 famine and, 60–61
Borrowing of concepts, 279–82 in France, 63
Botiza (Romania) glacial ice-core data and, 49
collectivisation and, 220 greenhouse gases and, 54–55
ecosystem and land-use diversity in, 215 in Greenland, 56
Boucher de Perthes, Jacques, 14 during Holocene Period, 42–46
Bourdieu, Pierre, 146 human relationships and, 58–67
Bourgois, P., 23 during Ice Ages, 42
Boyd, R., 153 in India, 51
Braided river, time as, 32–33 Late Antique Little Ice Age, 47, 56
Braudel, Fernand, 18–19 Little Ice Age, 51–54, 57, 58
Buffering, biological diversity and, 185–86 Maya civilisation and, 56
Bull, J.W., 116 Medieval Warm Period, 50–51, 56, 57, 58
Burgundy (France) Mid-Holocene Thermal Maximum, 43
droughts in, 279 migration and, 60, 64
ponds in, 260–61 neoglaciation, 43–45
Butzer, Karl, 147 polar amplification, 46
proxy data for, 47–49
Canada, pristine environments in, 118 religious persecution and, 63–64
Cascade effects, 303 Roman Warm Period, 46–47
Çatalhöyük (Turkey), domesticated sediment archives and, 49
landscapes in, 166–67 societal impact of, 55–59
Causation in complex adaptive systems, societal response to, 58–67
24, 275–76 speleothem records and, 49
Central Asia, climate change in, 56 Suess cycle, 45–46
Centre for Biodiversity (CBM), 3 Climax communities, 150
China Coalitions, 310–11
climate change in, 46–47, 51, 56, 57–58 Collaborative research, 7–8, 9, 240–42,
scientific traditions in, 17 265–66, 267–68
Circumpolar Networks Coalition, 311 authorship and, 252–55
Clements, F.E., 22, 150 ‘bad room’, 249–50
Climate change, 6, 41–42, 67–68 benefits of, 240–42
adaptation strategies, 61–63 checklist, 266–67
agriculture and, 60–61 common research questions, 242–43
Akkadian Empire and, 56 communication in, 248–49
Anasazi people and, 56 crowdsourced funding of, 256–57
archives and, 306–07 egalitarian nature of, 249
Arctic amplification, 46 ‘epistemological Esperanto’ and, 251, 281
armed conflict and, 57–58, 65–66 ‘followership’ in, 250–51

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:54:22, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
8
1
3

318 Index

Collaborative research (cont.) conceptual framework of historical


funding of, 255–58 ecology
‘good room’, 249–50 Conjoctures (contexts), 18
heterarchy in, 250 Conservation
integration of data and methods baselines in, 114–16
in, 258–63 biocultural awareness and, 116–20
integration-oriented cooperation and biocultural diversity and, 304
boundary-crossing, 282 essential components of past,
interdisciplinary research, 242, 282–83 identifying, 290–91
language and, 251 historical awareness and, 6–7, 116–20
leadership in, 249–50 historical ecology approaches in, 287–89
learning and, 252 human-nature interactions in, 287–89
local inhabitants, involvement of, 243–45 multi-dimensional past and, 289–90
management of data in, 263–65 of threatened species, 289
matrix management in, 250 Continuity, biological diversity and, 184
multidisciplinary research, 242 Contributors, 3–4
necessity of, 240–42 Convention on Biological Diversity (CBD), 3
new sources of knowledge, 283–85 baselines and, 115
in practice, 285–87 biological diversity defined in, 183
publication and, 252–55 traditional ecological knowledge (TEK) in
stakeholder involvement in, 243–47 rural communities and, 104–05
transdisciplinary research, 242–43, Cowles, H.C., 150
245–48, 282–83 Crop diversity among yeomen in colonial
Collective memory versus individual America, 201–03
memory, 127–29 Crowdsourced funding of collaborative
Colonial America, yeomen in. See Yeomen research, 256–57
in colonial America ‘Cultural amnesia’, 124
Common Agricultural Policy (CAP), 103–04 Cultural diversity, 7, 186–87See also
Common-pool resources (CPR), 207–08 Biocultural diversity
Communication in collaborative among yeomen in colonial America,
research, 248–49 199–206 (See also Yeomen in colonial
Communities of practice (CoP), America)
106–07, 301–02 differences with biological
Competition for resources, biological diversity, 188–89
diversity and, 185, 190 differing perspectives on, 191–92
Complex adaptive systems in grassland-based agriculture in Romania
advantages of in historical ecology, 30–31 (See Grassland-based agriculture in
causation in, 24, 275–76 Romania)
connectivity in, 25 human-nature relationships,
context in, 26 understanding mechanisms of,
defined, 22 192–93
emergence in, 26 interaction with biological
features of, 22–23 diversity, 189–91
history in, 25–26 preservation of, 287–91
investigation of time in, 21–25 in reindeer husbandry in Sweden (See
properties of, 25–27 Reindeer husbandry in Sweden)
terminology in, 22 similarities with biological
time in, 26–30 diversity, 189–91
tipping points in, 26–29, 302–03 in summer farming in Sweden (See
Concepts for integrated research in Summer farming in Sweden)
historical ecology. See Integrative ‘Cultures of catastrophe’, 276

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:54:22, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
9
1
3

Index 319

Dallimer, M., 124 glades as, 157–59, 161–62


Dark Age Cooling Period, 47, 56 grasslands as, 162–66
Darwin, Charles, 16, 152 integrative conceptual framework
Data in collaborative research and, 155–69
integration of, 258–63 livestock pens and, 159–60, 167–69
management of, 263–65 in Turkey, 166–67
Dataverse (database), 265 umbrella thorn ands, 160–61
Daw, T.M., 124 Domestication, 280–81
The Decline of the West (Spengler), 17 Drew, J., 124
Definition of historical ecology, 1–2
Definitions, going beyond, 280–81 Ecological age, biological diversity and, 184
DeLanda, M., 20 Ecological mutualism, 160–61
Deleuze, G., 148 Ecology, time in, 16
Dendroclimatology, 48 Egalitarian nature of collaborative
Denevan, W.M., 118 research, 249
Dependence Einstein, Albert, 20–21
in archaeology, 149–50 Ellis, E.C., 117–18, 186–87, 189
domesticated landscapes and, 167–69 Emic time, 19–20
DeVore, I., 153 Engels, Friedrich, 146–47
De Vries, D., 121 England
de Vries cycle, 45–46 agriculture in, 228
Dialectics, 146–47 climate change in, 63
Differences, taking advantage of, 281–82 Poor Law of 1602, 63
Distributed Long Term Observing Network sheep farming in, 124
of the Past, 307 Enlightenment perspective on ecology, 300
Diversity, 7, 182–83 Entanglement
among yeomen in colonial America, in archaeology, 149–50
199–206 (See also Yeomen in colonial biocultural diversity and, 304–06
America) domesticated landscapes and, 166–69
biocultural diversity (See Biocultural quantum entanglement, 30–31
diversity) Environmental determinism, 57
biological diversity (See Biological ‘Epistemological Esperanto’, 251, 281
diversity) Erdal, D., 153
cultural diversity (See Cultural diversity) Ethnic diversity, 192
ethnic diversity, 192 Europe. See also specific country
in grassland-based agriculture in Romania domesticated landscapes in, 162–66
(See Grassland-based agriculture in grasslands in, 162–66
Romania) European Union
linguistic diversity, 186, 192 Common Agricultural Policy
of livelihood, 186–87, 189 (CAP), 103–04
in reindeer husbandry in Sweden (See Mediterranean Mountainous Landscapes
Reindeer husbandry in Sweden) in Europe Project, 304
in summer farming in Sweden (See Water Framework Directive
Summer farming in Sweden) (WFD), 121–22
terminology in, 183 Événements (events), 18
Domesticated landscapes, 155–56 Evolutionary age, biological diversity
in Africa, 156–62 and, 184
dependence and, 167–69 Evolution of historical ecology, 300–01
ecological mutualism in, 160–61 Exaltation of the Holy Cross, 96–97
entanglement and, 166–69 Excursions, 278
in Europe, 162–66 ‘Extinction of experience’, 134–35

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:54:22, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
0
2
3

320 Index

Famine, climate change and, 60–61 Goal of historical ecology, 2


Feast of St. Anne, 98 Godemann, J., 282
Feast of St. Demetrius, 97 Gómez-Baggethun, Erik, 251n. 1
Feast of St. Elijah, 98 ‘Good room’, 249–50
Feast of St. Foca, 98 Gould, S.J., 20
Feast of St. John, 96–97 Grassland-based agriculture in
Feast of St. Mary Magdalene, 98 Romania, 215
Feast of St. Michael, 96 biocultural diversity and, 221–22
Feast of St. Pantaleon, 98 co-evolution of practices in, 218–19
Feasts, 96–97 ecosystem and land-use diversity, 215
Febvre, Lucien, 17 evolution of diversity in, 218–19
Federation of Swedish Farmers, 257–58 flexibility in, 217–18
Ferraro, P.J., 116 historical ecology and, 227
Finland historical perspectives on, 219–21
disposal of nuclear fuel in, 130 meadows and, 217–18
reindeer husbandry in Sweden and, 25–26 scales of diversity in, 215–17
‘Followership’ in collaborative spatial and temporal complexity
research, 250–51 and, 224–25
Food availability, biological diversity types of diversity in, 215–17
and, 184 variation in, 217–18
Forecasting, 279 Grasslands
Forestry adaptation and, 165
management of resources in, 303–04 as domesticated landscapes, 162–66
reindeer husbandry in Sweden grassland-based agriculture in Romania
and, 195–96 (See Grassland-based agriculture in
Frames of reference. See Baselines Romania)
France niche construction and, 165–66
agriculture in, 227–28 Greenhouse gases, 54–55
climate change in, 63 Greenland, climate change in, 56
droughts in, 279 Guattari, F., 148
ponds, study of, 260–61
Frere, John, 14 ‘Habitus’, 146
Future Earth (NGO), 310 HagmarksMISTRA Programme,
246–47, 257–58
Gamla Uppsala (Sweden), mounds at, 14 Haila, Y., 120
Gavriely-Nuri, D., 127 Hardin, Garrett, 207
Genetic material, biological diversity Hegel, G.W.F., 146–47
and, 284 Henrich, J., 153
Geology Heterarchy in collaborative research, 250
communication in, 248 Hicks, Megan, 308
time in, 13–14, 31–32 Higgs, E., 170
Giddens, Anthony, 19, 146 Histoire de climat depuis l'an mil
GIS software, 258 (Ladurie), 19
Glacial ice-core data, 49 Historical ecology. See specific topic
Glades, 157–59, 161–62 Hodder, I., 149–50, 166–68
Gladwell, Malcolm, 26–28 Hoffman, S.M., 276
Gleason, H.A., 150 Holistic theories, 150
Global Human Ecodynamics Alliance Holling, C.S., 16
(GHEA), 311 Holocene Period, climate change
Global warming. See Climate change during, 42–46

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:54:22, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
1
2
3

Index 321

Homer-Dixon, T., 29–30 International Union for Conservation of


Horstkotte, T., 199 Nature (IUCN), 287
Hunter, M., 116–17 Inter-Tropical Convergence Zone (ITCZ),
Hutchinson, G.E., 152 43, 51, 53, 67
Hutton, James, 14 Iron Age, 165
Islamic State of Iraq and the Levant (ISIS/
Ice Ages, 42 ISIL), 32
Iceland, traditional ecological knowledge Italy, biocultural heritage in, 304
(TEK) in, 308
India Jackson, J.B.C., 116–17
climate change in, 51 Jarvis, Jon, 305–06
scientific traditions in, 17 Johnson, N., 25
‘Individualistic concept of ecology’, 150–51
Individual memory versus collective Kimmerer, Robin Wall, 306
memory, 127–29 Knowlton, N., 116–17
Innovation, 307–08
Integrated History and Future of People on Labour
Earth (IHOPE), 311 agriculture, work exchanges and, 94–95
Integration of data and methods in yeomen in colonial America and, 200–01
collaborative research, 258–63 Ladurie, Emmanuel Le Roy, 19
Integration-oriented cooperation and Lake Rotehogstjärnen (Sweden), traditional
boundary-crossing, 282 ecological knowledge (TEK) in, 122
Integrative conceptual framework of Landscape ecology, 147–48
historical ecology, 7, 145–46, 169–70 Landscapes
adaptation, revisiting, 153–55 biological diversity and, 284
changing concepts and, 148–51 defined, 282
dialectics and, 146–47 domesticated landscapes (See
domesticated landscapes and, 155–69 (See Domesticated landscapes)
also Domesticated landscapes) ‘mosaic landscapes’, 150–51
entanglement and, 149–50 Lange, H., 25
historical background, 146–48 Language, collaborative research and, 251
history and, 169–70 Last Glacial Maximum, 42
holistic theories, 150 Late Antique Little Ice Age, 47, 56
‘individualistic concept of ecology’, Law of the Few, 27–28
150–51 Leadership in collaborative research, 249–50
networks of knowledge, 148–49 Learning, collaborative research and, 252
new framework for understanding socio- Levin, S.A., 22
environmental relations, 151 Levy, D., 126
niche construction, revisiting, 151–53 Libby, Walter, 15
Interdisciplinary research, 242, 282–83 Linguistic diversity, 186, 192
Intergovernmental Panel on Climate Linnaeus, Carl, 16
Change (IPCC), 55, 115 Little Ice Age, 51–54, 57, 58
International Council for Science (ICSU), Livelihood, diversification of, 186–87, 189
309–10 Livestock
International Geosphere-Biosphere Africa, livestock pens in, 159–60, 167–69
Programme, 309n. 1 yeomen in colonial America, livestock
International Human Dimensions diversity among, 204–06
Programme, 309n. 1 Longues durées (long-term trends), 18
International Social Science Council Lucas, G., 20
(ISSC), 309–10 Lyell, Charles, 14–15

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:54:22, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
2
3

322 Index

Maffi, L., 187–88 Native Americans, 289–90, 306


Management of data in collaborative Natural resources
research, 263–65 agriculture, common resources and, 92–94
Management of resources, 303–04 biological diversity, competition for
Maramureș (Romania) resources and, 185, 190
collectivisation and, 220 common-pool resources (CPR), 207–08
ecosystem and land-use diversity in, 215 management of resources, 303–04
milk-measuring feasts in, 92–93 reindeer husbandry in Sweden, diversity
mowing in, 96 of resources and, 198
Marine sciences, baselines in, 114–15 sustainability of resources, 292–94 (See
Maron, M., 114–15 also Sustainability of resources)
Marx, Karl, 146–47 traditional ecological knowledge (TEK),
Matrix management in collaborative common resources and, 92–94
research, 250 Neoglaciation, 43–45
Maurer, K., 190 Nested logical types, 21
Maya civilisation Networks of knowledge, 148–49
climate change and, 56 New Archaeology, 15, 22–23
water management in, 308 New sources of knowledge, 283–85
McCloskey, D.N., 227–28 Newton, Isaac, 20–21
McDonald-Madden, E., 116 Niche construction
McGlade, J., 19, 32 grasslands and, 165–66
Meadows, D.H., 23 revisiting, 151–53
Medieval Warm Period, 50–51, 56, 57, 58 Nongovernmental organisations
Medin, D., 134–35 (NGOs), 309–10
Mediterranean Mountainous Landscapes in Nora, P., 130
Europe Project, 304 Nordic Network for Interdisciplinary
La Méditerranée et le Monde Méditerranéen Environmental Studies (NIES), 311
a l'époque de Philippe II (Braudel), 18 North, Douglass C., 206
Memory North Atlantic Biocultural Organization
‘cultural amnesia’, 124 (NABO), 311
implications of research, 131–34 Noss, N., 134–35
individual versus collective Nuclear fuel, disposal of, 130
memory, 127–29 Nuer people, 19
mnemonic practices, 129–31 Nuttall, M., 130–31
‘personal amnesia’, 123–24 Nykvist, B., 126
Shifting Baseline Syndrome (SBS)
and, 125–27 Odling-Smee, F.J., 152
‘social-ecological memory’, 129 Odum Institute, 265
as ‘something people do’, 125–27 Olick, J.K., 126
traditional ecological knowledge (TEK) Oliver-Smith, A., 276
and, 292 Olof Rudbeck the Elder (Sweden), 14
Methods in collaborative research, On the Origin of Species (Darwin),
integration of, 258–63 16, 152
Mid-Holocene Thermal Maximum, 43 Operationalisation of collaborative research.
Migration, climate change and, 60, 64 See Collaborative research
Minos (Greek mythology), 2 Ostrom, Elinor, 206
Mnemonic practices, 129–31
Moen, J., 199 Papworth, S.K., 123–24
‘Mosaic landscapes’, 150–51 Parrott, L., 25
Mowing, 96 Pastoral Iron Age, 161
Multidisciplinary research, 242 Pastoral Neolithic Period, 161

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:54:22, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
3
2

Index 323

Pastures temporal dynamics and, 197–98


biocultural diversity and, 224–25 winter forage and, 194–95
grassland-based agriculture in Romania Religious persecution, climate change
and (See Grassland-based agriculture in and, 63–64
Romania) Research
reindeer husbandry in Sweden and, 198 collaborative research (See Collaborative
summer farming in Sweden and (See research)
Summer farming in Sweden) concepts for integrated research (See
Pattanysak, S.K., 116 Integrative conceptual framework of
Pauly, D., 123 historical ecology)
Pennsylvania, yeomen in. See Yeomen in historical ecology as research
colonial America framework, 2
‘Personal amnesia’, 123–24 interdisciplinary research, 242,
Perz, S.G., 303–04 282–83
Peter (Saint), 99 multidisciplinary research, 242
Piaget, Jean, 267 taking into action, 8, 309–11
Pielou, Evelyn, 23 taking research into action (See Action,
Pinker, S., 153 taking research into)
Polar amplification, 46 transdisciplinary research, 242–43,
Populations, biological diversity and, 284 245–48, 282–83
Power of Context, 27–28 Resilience, biological diversity and,
Practice and theory, historical ecology in. 185, 189
See Theory and practice, historical Resources. See Natural resources
ecology in Rewilding, 191–92
Predation, biological diversity and, 185 Richerson. P.J., 153
Predictability, biological diversity and, 184 Risk management in agriculture,
Principles of Geology (Lyell), 14–15 227–28
Pristine environments, 116–20 Roediger, H.L., 126
‘Pristine myth’, 118 Romania
Proxy data agriculture in, 102–03
for climate change, 47–49 biological diversity in, 105–06
excursions, 278 collectivisation in, 88–89, 220
Publication, collaborative research common resources in, 92–93
and, 252–55 de-collectivisation in, 89–90
feasts in, 96–97
Quantum entanglement, 30–31 grassland-based agriculture in (See
Quaternary Period, 42 Grassland-based agriculture in
Romania)
Radiometric dating, 15 milk-measuring feasts in, 92–93
Reason, historical ecology and, 300 mowing in, 96
Reindeer husbandry in Sweden, 193–95 pastures in, 92–93
biocultural diversity and, 221–22 seasonality and timing in agriculture
diversity of resources and, 198 in, 95–97
forestry and, 195–96 traditional ecological knowledge (TEK) in,
history, impact of, 25–26 88–90, 102–04
knowledge of diversified land use villages in, 89
and, 198–99 weather prediction in, 97–100
mismatch of scales and, 196–97 work exchanges in, 94
pastures and, 198 Roman Warm Period, 46–47
spatial and temporal complexity and, 225 Roturier, S., 198–99
spatial variation and, 197–98 Roué, M., 198–99

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:54:22, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
4
2
3

324 Index

Rural communities, traditional ecological Semi-natural biotopes, 163


knowledge (TEK) in, 6, 84–85 Sendai Framework for Disaster Risk
agriculture and, 85, 91–92 Reduction, 276
biological diversity and, 105–06 Sheep farming, 124
in changing world, 88–91 Shifting Baseline Syndrome (SBS), 6–7,
Common Agricultural Policy (CAP) 112–14, 122–25, 134–35
and, 103–04 ‘cultural amnesia’ and, 124
common resources and, 92–94 defined, 113–14, 123
communities of practice (CoP) implications of research, 131–34
and, 106–07 individual versus collective
comparative analysis, 99–100 memory, 127–29
de-valuation of, 103–08 memory and, 125–27
disruption of, 88 mnemonic practices, 129–31
dynamics of, 85–88 ‘personal amnesia’ and, 123–24
evolution of, 85–88 ‘social-ecological memory’ and, 129
feasts and, 96–97 Smith, William ‘Strata’, 14
in Iceland, 308 Social constructionism, 113n. 1
loss of, 89, 91 ‘Social-ecological memory’, 129
modernisation and, 91 Social nature of humanity, 301
mowing and, 96 Soranno, P.A., 29
path dependency and, 87–88 Spain, biocultural heritage in, 304
recognition of, 104–05 Spatial heterogeneity, biological diversity
revaluation of, 103–08 and, 185, 190
in Romania, 88–90, 102–04 Spatial transference of historical ecology
seasonality and timing in agriculture insights, 276–78
and, 95–97 Species richness, 184, 185
in Sweden, 90–91, 101–02, 103, 122 Speleothem records, 49
terminology in, 85 Spengler, Oswald, 17
weather prediction and, 97–100 Spiritual values, 306
work exchanges and, 94–95 Stability, biological diversity and,
Russell, Bertrand, 21 184, 185–86
Russia, reindeer husbandry in Sweden Steps to an Ecology of Mind (Bateson), 21
and, 25–26 Stickiness Factor, 27–28
Stommel, Henry, 29
St. George’s Day, 97 Stratigraphy, 14
Sami people, reindeer husbandry and. See Structuration, 146
Reindeer husbandry in Sweden Suess cycle, 45–46
SBS. See Shifting Baseline Syndrome (SBS) Summer farming in Sweden, 206–08
Scales adaptation and, 226
grassland-based agriculture in Romania, biocultural diversity and, 221–22
scales of diversity in, 215–17 common-pool resources (CPR)
reindeer husbandry in Sweden, mismatch and, 207–08
of scales and, 196–97 decline of, 214
sustainability of resources, conflict among dispute resolution and, 213–14
scales in, 293–94 evolution of, 208–10
Science, historical ecology and, 300 institutions and, 206–08
Seasonality monitoring, 213
agriculture and, 95–97 resource areas, defining, 211–12
biological diversity and, 184–85, 190 rules, creation of, 212–13
Sediment archives, 49 spatial and temporal complexity and, 225

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:54:22, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
5
2
3

Index 325

‘tragedy of the commons’ and, 207 Theory and practice, historical ecology
use restrictions adapted to local in, 8, 275
conditions, 212 biological diversity, preservation
‘Superorganisms’, 150 of, 287–91
Șurdești (Romania), weather prediction in, borrowing of concepts, 279–82
97, 98–99 causation in complex adaptive systems,
Sustainability of resources, 292–94 24, 275–76
conflict among scales in, 293–94 collaborative research (See Collaborative
human-nature interactions and, 293 research)
local versus societal sustainability, conservation, historical ecology
292–93 approaches in, 287–89
Svallfors, Stefan, 249–50 cultural diversity, preservation of, 287–91
Sweden definitions, going beyond, 280–81
agriculture in, 101–02 differences, taking advantage of, 281–82
common resources in, 93–94 interdisciplinary research, 282–83
feasts in, 96 new sources of knowledge, 283–85
mowing in, 96 spatial transference of historical ecology
nuclear fuel, disposal of, 130 insights, 276–78
pastures in, 90–91, 93–94 sustainability of resources and, 292–94
pristine environments in, 118–19 (See also Sustainability of resources)
Reindeer Husbandry Act, 193 temporal transference of historical
reindeer husbandry in (See Reindeer ecology insights, 278–79
husbandry in Sweden) traditional ecological knowledge (TEK)
seasonality and timing in agriculture and, 291–92
in, 95–97 transdisciplinary research, 282–83
summer farming in (See Summer farming Theseus (Greek mythology), 2
in Sweden) A Thousand Years of Non-Linear History
traditional ecological knowledge (TEK) in, (DeLanda), 20
90–91, 101–02, 103, 122 Threatened species, conservation of, 289
weather prediction in, 97–100 Time, historical ecology and, 6, 13
work exchanges in, 94–95 Annalistes and, 17–19
Swedish University of Agricultural in archaeology, 14–15
Sciences, 3 backcasting, 279
Swedish World Wide Fund for in biology, 16
Nature, 257–58 as braided river, 32–33
Switzerland in complex adaptive systems, 26–30
biocultural diversity in, 190 conjoctures (contexts), 18
terracing, 308 in ecology, 16
emic time, 19–20
Tanzania, conservation in, 121 événements (events), 18
‘Technology solution’, 300–01 excursions, 278
TEK. See Traditional ecological forecasting, 279
knowledge (TEK) in geology, 13–14, 31–32
Temperature, biological diversity and, 184 investigation of in complex adaptive
Terracing, 308 systems, 21–25
Themes longues durées (long-term trends), 18
constructive approaches, 5 nested logical types and, 21
dialogues, 5 new frameworks for, 31–32
long-term history of biological Newtonian versus Einsteinian
diversity, 5 models, 20–21

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:54:22, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780
6
2
3

326 Index

Time, historical ecology and (cont.) Universal Declaration on Cultural


search for new concepts of, 17–21 Diversity (UNESCO), 183, 187
temporal transference of historical United States
ecology insights, 278–79 National Parks Service, 304–06
tipping points, 26–29, 302–03 Native Americans, 289–90, 306
Tipping points, 26–29, 302–03 White Mountain Apache Tribe, 306
Tjeggelvas (Swedish nature reserve), pristine yeomen in colonial America (See Yeomen
environments in, 118–19 in colonial America)
Tooby, J., 153 Uppsala University, 3
Traditional ecological knowledge (TEK)
aesthetic and spiritual values and, 306 Valinia, S., 122
communities of practice (CoP), 301–02 Vegetation, biological diversity and, 284
in Iceland, 308 Vinitzky-Seroussi, V., 126
memory and, 292 von Heland, J., 126
in policy planning, 292 von Hofmannsthal, Hugo, 127
in Romania, 88–90, 102–04
in rural communities (See Rural Warming, E., 150
communities, traditional ecological Water availability, biological diversity
knowledge (TEK) in) and, 184
in Sweden, 90–91, 101–02, 103, 122 Water Framework Directive (WFD), 121–22
in theory and practice, 291–92 Water management, 308
‘Tragedy of the commons’, 207 Weather prediction, 97–100
Traits, biological diversity and, 284 Whitehead, A.N., 21
Transdisciplinary research, 242–43, White Mountain Apache Tribe, 306
245–48, 282–83 Whiten, A., 153
Tulving, E., 126 Winter forage, reindeer husbandry in
Turkey, domesticated landscapes in, 166–67 Sweden and, 194–95
Tylor, E.B., 186 Woodley, E., 187–88
Woody plants, biological diversity and, 284
Umbrella thorn, 160–61 Workshops, 3, 5–6
UNESCO World Climate Research Programme,
Convention for the Safeguarding of the 309n. 1
Intangible Cultural Heritage, 104 World Intellectual Property Organization, 85
International Council for Science and, 310
Universal Declaration on Cultural Yeomen in colonial America, 199–200
Diversity, 183, 187 adaptation and, 226
Uniformitarianism, 14–15 biocultural diversity and, 221–22
United Nations crop diversity and, 201–03
conferences, 309 historical ecology and, 227
Convention for the Safeguarding of labour and, 200–01
the Intangible Cultural Heritage livestock diversity and, 204–06
(UNESCO), 104 spatial and temporal complexity
Environmental Programme (UNEP), 287 and, 225
Sendai Framework for Disaster Risk
Reduction, 276 Zeuner, F.E., 155–56

Downloaded from https://www.cambridge.org/core. Access paid by the UCSB Libraries, on 02 Aug 2018 at 17:54:22, subject to the Cambridge
Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/9781108355780

You might also like