0% found this document useful (0 votes)
14 views302 pages

Historical Ecology - 2022

Uploaded by

renata ruiz
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
14 views302 pages

Historical Ecology - 2022

Uploaded by

renata ruiz
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 302

Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024].

See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Ecology
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Ecology

Present and Forecast the Future of Ecosystems

Guillaume Decocq
Learning from the Past to Understand the

Coordinated by
Field Directors – Françoise Gaill and Dominique Joly
Historical Ecology, Subject Head – Dominique Joly
Ecosystems and Environment
SCIENCES
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
First published 2022 in Great Britain and the United States by ISTE Ltd and John Wiley & Sons, Inc.

Apart from any fair dealing for the purposes of research or private study, or criticism or review, as
permitted under the Copyright, Designs and Patents Act 1988, this publication may only be reproduced,
stored or transmitted, in any form or by any means, with the prior permission in writing of the publishers,
or in the case of reprographic reproduction in accordance with the terms and licenses issued by the
CLA. Enquiries concerning reproduction outside these terms should be sent to the publishers at the
undermentioned address:

ISTE Ltd John Wiley & Sons, Inc.


27-37 St George’s Road 111 River Street
London SW19 4EU Hoboken, NJ 07030
UK USA

www.iste.co.uk www.wiley.com

© ISTE Ltd 2022


The rights of Guillaume Decocq to be identified as the author of this work have been asserted by him in
accordance with the Copyright, Designs and Patents Act 1988.

Any opinions, findings, and conclusions or recommendations expressed in this material are those of the
author(s), contributor(s) or editor(s) and do not necessarily reflect the views of ISTE Group.

Library of Congress Control Number: 2022940278

British Library Cataloguing-in-Publication Data


A CIP record for this book is available from the British Library
ISBN 978-1-78945-090-3

ERC code:
PE10 Earth System Science
PE10_3 Climatology and climate change
PE10_4 Terrestrial ecology, land cover change
LS8 Ecology, Evolution and Environmental Biology
LS8_4 Evolutionary ecology
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Contents
Chapter 1. A General Introduction to Historical Ecology . . . . . . . . . . . . 1
Guillaume DECOCQ

1.1. The roots of historical ecology . . . . . . . . . . . . . . . . . . . . . . . . . . . 2


1.2. A multidisciplinary approach of socio-ecosystems . . . . . . . . . . . . . . . . 3
1.3. Recent trends in historical ecology . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4. The way forward . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.5. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

Chapter 2. Historical Resurveys Reveal Causes of Long-term


Ecological Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
Donald M. WALLER

2.1. Serious ecological changes are pervasive . . . . . . . . . . . . . . . . . . . . . 11


2.2. Anthropogenic drivers of ecological change . . . . . . . . . . . . . . . . . . . 12
2.2.1. The missing baseline problem . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2.2. Ecological communities are complex . . . . . . . . . . . . . . . . . . . . . 13
2.3. Kinds of ecological change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.3.1. Natural community dynamics . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.3.2. Anthropogenic drivers of ecological change . . . . . . . . . . . . . . . . . 14
2.4. Understanding the forces driving ecological change . . . . . . . . . . . . . . . 18
2.4.1. Natural experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.4.2. Metrics of change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.4.3. Can functional traits reveal drivers of change? . . . . . . . . . . . . . . . . 19
2.4.4. Vectors of change – ordination . . . . . . . . . . . . . . . . . . . . . . . . 20
2.5. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.6. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
vi Historical Ecology

Chapter 3. Getting the Right Answer Can Take a While: Long-term


Ecological Field Studies as Historical Ecology . . . . . . . . . . . . . . . . . . 27
Frank S. GILLIAM

3.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.2. Fernow Experimental Forest . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.2.1. Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.2.2. Site description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.2.3. Field design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.3. Long-term studies at Fernow Experimental Forest, West Virginia . . . . . . . 32
3.3.1. Effects of acidification on soil fertility and herb layer
cover and foliar nutrients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.3.2. Effects of N addition on soil N dynamics . . . . . . . . . . . . . . . . . . . 34
3.3.3. Effects of N addition on herb layer composition and diversity . . . . . . . 35
3.3.4. The N homogeneity hypothesis . . . . . . . . . . . . . . . . . . . . . . . . 37
3.3.5. A look at the future: declines in the atmospheric deposition of N . . . . . 40
3.4. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.5. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

Chapter 4. Gaps and Cracks in Land Cover Mapping for Historical Ecology 45
Francesca Di PIETRO, Roger COLY, Clémence CHAUDRON, Samuel LETURCQ

4.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.2. Three main steps of past land cover mapping . . . . . . . . . . . . . . . . . . . 46
4.3. Land cover in the 19th century: the old cadasters . . . . . . . . . . . . . . . . . 47
4.4. Land cover in the 20th century: aerial photographs . . . . . . . . . . . . . . . . 50
4.5. Present land cover: modern databases . . . . . . . . . . . . . . . . . . . . . . . 52
4.6. From different sources to one land cover typology . . . . . . . . . . . . . . . . 53
4.7. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
4.8. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

Chapter 5. The Use of Repeat Photography in African Historical Ecology 57


Michael Timm HOFFMAN and Rick F. ROHDE

5.1. Repeat photography as an emerging tool in African historical ecology . . . . . 57


5.2. Repeat photography and landscape change in Africa . . . . . . . . . . . . . . . 58
5.2.1. Early contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.2.2. Ethiopia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.2.3. Southern Africa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.3. Long-term change in plant populations as revealed by repeat photography . . 62
5.4. Strengths and limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.5. Future directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
5.6. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Contents vii

Chapter 6. Remote Sensing for Historical Ecology. . . . . . . . . . . . . . . . 71


Pierre-Alexis HERRAULT and David SHEEREN

6.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
6.2. Landscape spatio-temporal changes as a proxy of biodiversity . . . . . . . . . 72
6.3. Mapping landscapes at different dates . . . . . . . . . . . . . . . . . . . . . . . 73
6.3.1. Airborne laser scanning data . . . . . . . . . . . . . . . . . . . . . . . . . . 73
6.3.2. Historical maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
6.3.3. Old aerial photographs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
6.3.4. Satellite images . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
6.4. Modeling the effects of spatio-temporal changes on present-day biodiversity . 77
6.4.1. Structural spatio-temporal metrics . . . . . . . . . . . . . . . . . . . . . . . 77
6.4.2. Functional spatio-temporal metrics . . . . . . . . . . . . . . . . . . . . . . 79
6.5. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

Chapter 7. Soil Archives: Where Soilscape History Meets Present-day


Ecosystems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
Boris BRASSEUR, Damien ERTLEN and Vincent ROBIN

7.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
7.2. Mechanisms of soil archiving and the associated dynamics . . . . . . . . . . . 86
7.2.1. Pedoturbations of biological and physical origins . . . . . . . . . . . . . . 86
7.2.2. Eluviation–Illuviation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
7.2.3. Anthropogenic factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
7.2.4. Effects of geomorphological processes on soil archives . . . . . . . . . . . 88
7.3. Examples of soil archives and their influence on current ecosystems . . . . . . 90
7.3.1. Chemical archives, witnesses of progressive soil transformations . . . . . 90
7.3.2. Physical archives: reading the soil pit profile and microtopographic
features . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
7.3.3. Soil organic matter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
7.3.4. Botanical remains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
7.4. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
7.5. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

Chapter 8. Continuous and Nested Time in Historical Ecology:


Application to Soil Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
Damien ERTLEN

8.1. Interdisciplinarity and time in historical ecology . . . . . . . . . . . . . . . . . 99


8.2. Continuous time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
8.3. Nested time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
8.4. Different disciplines, different tools . . . . . . . . . . . . . . . . . . . . . . . . 103
8.5. Examples of nested and continuous time: soils and strata . . . . . . . . . . . . 105
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
viii Historical Ecology

8.6. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107


8.7. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

Chapter 9. The Analysis of Relic Charcoal Kilns for the Assessment of


Forest Trajectories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
Vincent ROBIN, Alexa DUFRAISSE and Claudia OLIVEIRA

9.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111


9.2. Looking at the platform of the kiln . . . . . . . . . . . . . . . . . . . . . . . . . 112
9.2.1. Looking at the dimensions of the kiln platforms . . . . . . . . . . . . . . . 112
9.2.2. Platform inventory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
9.3. Looking at the charcoal pieces . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
9.3.1. Sampling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
9.3.2. Taxonomic identification . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
9.3.3. Dendro-anthracology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
9.4. Looking at the ages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
9.5. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
9.6. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

Chapter 10. Ancient Trees and Botanical Indicators as Evidence for


Change and Continuity in Landscape Evolution . . . . . . . . . . . . . . . . . 123
Ian D. ROTHERHAM

10.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123


10.2. What is ancient woodland? Questions of woods versus old-growth forest,
and of continuity versus antiquity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
10.3. The value of ancient woods . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
10.4. Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
10.4.1. Evidencing ancient woodlands and the use of indicators. . . . . . . . . . 125
10.4.2. Tree form and growth as evidence of antiquity and continuity . . . . . . 128
10.4.3. The importance of ancient and veteran trees in woodland . . . . . . . . . 129
10.4.4. Soils and sediments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
10.5. An emerging woodland paradigm. . . . . . . . . . . . . . . . . . . . . . . . . 131
10.6. A simple new conceptual framework . . . . . . . . . . . . . . . . . . . . . . . 131
10.7. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
10.8. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133

Chapter 11. Towards a Methodological Framework for Investigating the


Hidden History of Woodland Covers . . . . . . . . . . . . . . . . . . . . . . . . . 135
Damien MARAGE, Catherine FRUCHART, Isabelle JOUFFROY-BAPICOT, Olivier GIRARDCLOS,
Vincent BALLAND

11.1. Why talk about hidden history when studying forest vegetation? . . . . . . . 135
11.2. From recent forests: a synecological point of view . . . . . . . . . . . . . . . 136
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Contents ix

11.3. From the walls: ancient documents and maps . . . . . . . . . . . . . . . . . . 136


11.4. From the wood: dendrochronology . . . . . . . . . . . . . . . . . . . . . . . . 139
11.5. From the ground: palynology . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
11.6. From the air: LiDAR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
11.7. Discussion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
11.8. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

Chapter 12. The Gate to the Forest is in its History . . . . . . . . . . . . . . . 151


Keith J. KIRBY

12.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151


12.2. The ancient woodland idea . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
12.3. Legacies of woodland management. . . . . . . . . . . . . . . . . . . . . . . . 153
12.4. Seeing the trees, not the woods . . . . . . . . . . . . . . . . . . . . . . . . . . 154
12.5. Exploring the distant past . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
12.6. Trees and woods from the past to the future . . . . . . . . . . . . . . . . . . . 157
12.7. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158

Chapter 13. Plant Assemblages and Ecosystem Functioning,


a Legacy of Long-term Interactions with Large Herbivores . . . . . . . . . . 163
Christophe BALTZINGER and Anders MÅRELL

13.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163


13.2. Large herbivores are ecosystem dominant interactors. . . . . . . . . . . . . . 164
13.2.1. Large herbivores as ecosystem engineers . . . . . . . . . . . . . . . . . . 164
13.2.2. Large herbivores and plant assemblages. . . . . . . . . . . . . . . . . . . 166
13.3. Long-term effects and methodological changes . . . . . . . . . . . . . . . . . 167
13.3.1. Paleoecological records . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
13.3.2. Modern data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
13.4. Plant–herbivore interactions over the long-term . . . . . . . . . . . . . . . . . 168
13.4.1. Quaternary communities of large herbivores and associated flora . . . . 168
13.4.2. The forest in the early Holocene . . . . . . . . . . . . . . . . . . . . . . . 169
13.5. Modern vegetation trajectories driven by large herbivores . . . . . . . . . . . 170
13.5.1. Herbivory effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
13.5.2. Temporal trajectories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
13.6. Perspectives, rewilding and ecosystem restoration . . . . . . . . . . . . . . . 172
13.7. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
x Historical Ecology

Chapter 14. A Historical Ecology of the Compiègne Forest (N France) . . 177


Jérôme BURIDANT, Boris BRASSEUR, Hélène HOREN, Emilie GALLET-MORON and
Guillaume DECOCQ

14.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177


14.2. The ancient forest: an intensively managed agricultural landscape? . . . . . . 178
14.3. The Medieval forest: a woodland (re)birth or a savanna-like ecosystem? . . . 184
14.4. The contemporary forest (19th century onward): a closed-canopy
multifunctional woodland . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
14.5. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
14.6. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191

Chapter 15. The Chestnut Orchards in the Bolognese Apennines:


A Vanishing Socio-ecological Habitat . . . . . . . . . . . . . . . . . . . . . . . . 195
Giovanna PEZZI, Fabrizio FERRETTI, Alberto MALTONI, Patrik KREBS, Marco CONEDERA
and Giorgio MARESI

15.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195


15.2. The traditional chestnut orchards . . . . . . . . . . . . . . . . . . . . . . . . . 197
15.3. The chestnut groves of the Bolognese Apennines . . . . . . . . . . . . . . . . 198
15.4. A changing world: abandonment, diseases and other problems . . . . . . . . 199
15.5. The turning point of the 1980s . . . . . . . . . . . . . . . . . . . . . . . . . . 199
15.6. Current constraints and future perspectives . . . . . . . . . . . . . . . . . . . 200
15.7. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203

Chapter 16. Claudius’ Coin in the Forest – Niche Construction and


Strategies by Early Colonizers of Boreal Inlands in Central
Scandinavia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
Ove ERIKSSON and Karl-Johan LINDHOLM

16.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207


16.2. Concepts and theoretical framework . . . . . . . . . . . . . . . . . . . . . . . 210
16.3. A historical overview of the colonization . . . . . . . . . . . . . . . . . . . . 211
16.4. A structured landscape . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
16.4.1. Constructing the environment . . . . . . . . . . . . . . . . . . . . . . . . 212
16.4.2. Managing livestock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
16.4.3. Shielings (secondary farms) . . . . . . . . . . . . . . . . . . . . . . . . . 214
16.5. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
16.6. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Contents xi

Chapter 17. Recent History of Vegetation Changes in the Arctic . . . . . . 221


Antoine BECKER-SCARPITTA, Bastien PARISY and Tomas ROSLIN

17.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221


17.2. The Arctic tundra biome . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
17.3. The Arctic historical ecological archive . . . . . . . . . . . . . . . . . . . . . 222
17.3.1. Remote sensing over time . . . . . . . . . . . . . . . . . . . . . . . . . . 223
17.3.2. Field-based records . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
17.4. Changes over time in tundra vegetation . . . . . . . . . . . . . . . . . . . . . 225
17.4.1. Changes in vegetation productivity . . . . . . . . . . . . . . . . . . . . . 225
17.4.2. Changes in vegetation phenology . . . . . . . . . . . . . . . . . . . . . . 226
17.4.3. Changes in plant community structure, composition and diversity . . . . 227
17.5. Synthesis and perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
17.6. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230

Chapter 18. Reconstructing the Impact of Humans on Aotearoa


New Zealand’s Biodiversity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
Nicolas J. RAWLENCE, Alexander J.F. VERRY, Karen GREIG, Justin J. MAXWELL,
Lara D. SHEPHERD and Richard WALTER

18.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233


18.2. Archaeological evidence for anthropogenic impact in New Zealand . . . . . 234
18.3. Paleovegetation change in pre- and post-European contact New Zealand. . . 237
18.4. Utilizing Aotearoa’s natural resources: Māori cultivation and translocation
of flora and fauna . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
18.5. Evolutionary consequences of Polynesian and European arrival . . . . . . . . 240
18.6. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
18.7. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243

Chapter 19. Historical Ecology of the Coastal Aeolian


Sedimentary Systems of the Canary Islands . . . . . . . . . . . . . . . . . . . . 247
Aarón Moisés SANTANA-CORDERO, Antonio Ignacio HERNÁNDEZ-CORDERO, Néstor
MARRERO-RODRÍGUEZ, Leví GARCÍA-ROMERO, Elisabet FERNÁNDEZ-CABRERA, Carolina
PEÑA-ALONSO, Emma PÉREZ-CHACÓN ESPINO and Luis HERNÁNDEZ-CALVENTO

19.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247


19.2. Study sites. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
19.3. Historical evolution of the coastal aeolian sedimentary systems of the
Canary Islands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
19.3.1. 19th century: territorial consolidation and spread of the agrarian
socioeconomic system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
19.3.2. 20th century to the present day: the tourism transformation . . . . . . . . 253
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
xii Historical Ecology

19.4. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255


19.5. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256

Chapter 20. Historical Forest Microclimates . . . . . . . . . . . . . . . . . . . . 259


Emiel DE LOMBAERDE, Karen DE PAUW, Pallieter DE SMEDT, Jonathan LENOIR, Camille
MEEUSSEN, Thomas VANNESTE, Kris VERHEYEN, Florian ZELLWEGER and Pieter DE FRENNE

20.1. Drivers of microclimate at the plot, forest and landscape scale . . . . . . . . 261
20.2. Methods to infer microclimate from the past and predict into the future . . . 265
20.3. Why do historical microclimates matter? Impacts on biodiversity from the
plot to landscape scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
20.4. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
20.5. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270

Chapter 21. Causes and Consequences of Extinction Debts:


Perspectives for Historical Ecology and Biological Conservation . . . . . . 273
Grégoire BLANCHARD and François MUNOZ

21.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273


21.2. Causes and processes entailing extinction debts . . . . . . . . . . . . . . . . . 274
21.3. Studying and detecting extinction debts from ecosystem history . . . . . . . 276
21.4. Implications for biodiversity conservation and management . . . . . . . . . . 280
21.5. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
21.6. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282

Chapter 22. Historical Ecology for the Past and the Future:
Organizing at Local and Regional Scales . . . . . . . . . . . . . . . . . . . . . . 285
Carole L. CRUMLEY

22.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285


22.2. Founding IHOPE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 286
22.3. Integrating the social sciences and humanities . . . . . . . . . . . . . . . . . . 287
22.4. Historical ecology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 288
22.5. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290
22.6. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 291

List of Authors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295

Index. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1

A General Introduction
to Historical Ecology
Guillaume DECOCQ
Jules Verne University of Picardie, Amiens, France

Since the second half of the 19th century, several human and social sciences
have increasingly incorporated the environment when interpreting human practices
and behavior. At the same time, ecological sciences have paid increasing attention to
human factors and history when analyzing biodiversity and functioning of forest
ecosystems and landscapes. Historical ecology emerged in this context.

How ecological systems have changed over time and how past human activities
impacted these dynamics are long-standing research questions in ecology. But the
emergence of historical ecology as a discipline aiming at answering these questions
is more recent. Though many definitions have been given in the literature, we can
simply define historical ecology, after Russel (1997), as the field of ecology which
aims to reconstruct the history of ecosystems and landscapes in order to analyze how
past events have impacted present days biodiversity and ecosystem functioning. This
includes causes and consequences of interactions between humans and the
environment (Beller et al. 2017), and as such, this definition is quite close to the one
used by anthropologists: an interdisciplinary research program concerned with
comprehending temporal and spatial dimensions in the relationships of human
societies to local environments and the cumulative global effects of these
relationships (Balée 2006). Historical ecology does not exclude any time period a
priori, even if most studies focused on historical times (i.e. for which written records
are available, even if the date varies among regions around the world).

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2 Historical Ecology

1.1. The roots of historical ecology

We can trace the roots of historical ecology back to several “undisciplined”


scientists, who paved non-conventional roads from their respective discipline,
independently from each other. An extensive review can be found in Szabó’s
epistemological essay (Szabó 2015). Seeking in the past the drivers of current
ecological patterns is an old idea. For example, in the 14th century, Dante already
described the historical degradation of Mediterranean forests. Later, in the
18th century, German foresters compared old records of forests to recent ones in
order to highlight the degradation of forest ecosystems (Stisser 1737), while in
England, Barrington published the first paper about indigenous versus introduced
tree species in 1769. Georges Perkins Marsh is usually considered as one of the
early pioneers of modern historical ecology, with the publication of his seminal
book “The Earth as Modified by Human Action” in 1874, though the term “historical
ecology” explicitly appeared in Tubbs’ book on the New Forest (1968).

Several disciplines have independently contributed to the emergence of historical


ecology:
– Forest history, sometimes labeled historical biogeography of forests. The
French historian Alfred Maury may be considered as the founder of this school, with
his remarkable thesis on the French forests from the pre-Roman to the modern times
(Maury 1850).
– Historical geography, especially through the French geographer Vidal de la
Blache, who launched the journal Annales de Géographie in 1891. This journal
published a large number of regional naturalist monographs focusing on landscapes
as human-made structures.
– Ecology and, more precisely, the field of paleoecology, which aims to
reconstruct past ecological communities and their dynamics in response to natural
and anthropogenic drivers using proxies such as pollens and macro-remains.
– Rural history, for which the French historians Fèbvre and Bloch were among
the pioneers via the journal Annales d’Histoire économique et sociale they launched
in 1929. This journal published a large number of papers dealing with impacts of the
environment on human activities and landscape planning.
– Landscape archaeology, with the seminal work of Hoskins in England,
published in 1955. This school was the first to use innovative tools such as aerial
photographs to detect and map archaeological artifacts at broader spatial scales, such
as former Roman villae in croplands. More recently, environmental archaeology
extended the scope of landscape archaeology to the reconstruction of the
relationships between past societies and the environments they lived in, using a wide
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
A General Introduction to Historical Ecology 3

range of proxies and covering the whole Quaternary era (Wilkinson and Stevens
2003).
– Environmental history, which roots in social science and ethnology, and aims
to study human interaction with the natural world over time. Interestingly,
environmental history emerged in North America with Nash (1967), and mostly
makes use of written sources and oral surveys.

In the second half of the 20th century, these various approaches progressively
converged and the old partition between disciplines tended to relax: time makes
historical ecology more and more a multidiscipline.

1.2. A multidisciplinary approach of socio-ecosystems

Current research projects in historical ecology most often rely on


multidisciplinary approaches, i.e. they involve already long-established disciplines
to focus on a joint topic or object. Most often, this object is a given ecosystem
or landscape, which is considered a socio-ecosystem (Redman et al. 2004).
Socio-ecosystems are hybrid systems: they are patterned by the interaction between
natural drivers (e.g. geomorphology, soil types, natural vegetation) and
anthropogenic drivers (e.g. landscape planning, land-use, resource exploitation). A
typical example is former orchards that experienced secondary succession following
abandonment (see Chapter 15). Investigating the history of socio-ecosystems thus
requires the analysis of two types of archives: ecological archives (e.g. soils, plant
remains) and human archives (e.g. written sources, archaeological remains).

Among ecological archives, soil is probably the most relevant one. Soils can
keep the memory of past environmental conditions for a very long time, and provide
historical ecologists with a number of physical, chemical and biological proxies
(see Chapter 7). Depending on their properties, they can preserve plant remains
(e.g. pollen, seeds and fruits, charcoals, phytoliths; e.g. Schoonmaker and Foster
1991; Feiss et al. 2017) and animal remains (Fitzpatrick and Keegan 2007), as well
as their DNA (Rawlence et al. 2014; Birks and Birks 2016; see also Chapter 18).
Even land-uses that took place a very long time ago left an imprint in soil properties,
which often still impact the current vegetation. For example, former Roman
settlements usually harbor a luxuriant vegetation dominated by eutrophic,
neutrophilous species, while on former Roman cultivated fields, the vegetation is
more scattered and dominated by acidophilous species (Dupouey et al. 2002; Plue et al.
2008; see also Chapter 14). Many case studies have been published for a range of
regions of Western Europe, all with similar results, so that some plant communities
are even used as indicators in field archaeological surveys (Decocq 2004) and can
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
4 Historical Ecology

guide targeted archaeological excavations. However, soil is a palimpsest and its


biological activity typically clouds the vertical layering of deposits, rendering
difficult the reconstruction of accurate time sequences, especially when a single
proxy is investigated (e.g. soil charcoals; Feiss et al. (2017), but see Chapter 9). It is
thus necessary to integrate several proxies with contrasted spatio-temporal resolution
to get accurate time reconstructions (see Chapter 8), and to cross-reference the
information with other ecological archives, such as botanical indicators (see
Chapters 10–12) and also human archives.

Regarding human archives, the availability of written sources is limited and both
their quantity and quality decrease as we go further back in the past. For example,
maps and cadasters allow us to reconstruct landscape changes over the last few
centuries, but with a low reliability for the most ancient ones (see Chapter 4). Aerial
photography allows us to cover only the last few decades but still with important
insights into fast-changing landscapes (see Chapter 5). Aside comparing the same
site at different times, remote sensing techniques can be used to seek archaeological
artifacts over extended areas, by evidencing vegetation (e.g. aerial photography) or
microtopographic (e.g. LiDAR) anomalies that the landscape has inherited from
past human activities (Challis et al. 2008; Beck 2011, see also Chapter 6).
Archaeological excavations greatly help in interpreting past human–ecosystem
interactions but are more rarely implemented in historical ecology projects, since
they can cover only restricted areas (but see Chapter 16).

1.3. Recent trends in historical ecology

Over the last few decades, studies in historical ecology have shifted from purely
qualitative and descriptive to more quantitative and mechanistic. Old concepts have
been revisited and novel insights have emerged. A typical example is the concept of
ancient woodlands. A number of descriptive studies, mostly from British plant
ecologists (Rackham 1980; Peterken 1981), evidenced differences in plant species
composition between ancient woodlands (i.e. woodlands that have continuously
existed since the date of the oldest available map, which usually dates back to the
end of the 18th century in most countries; Hermy and Verheyen 2007) and recent
woodlands (i.e. woodlands that have established on former agricultural lands, within
the last two centuries). Today, the concept of ancient forest species has largely
permeated vegetation science and conservation biology (see Chapters 10 and 12).
Now, historical ecology turns to understand the underlying mechanisms.
Experimental studies, mostly conducted in Belgium for Europe and at the Harvard
forest for North America, counterintuitively revealed that these compositional
differences were primarily explained by species’ dispersal capacities, rather than by
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
A General Introduction to Historical Ecology 5

habitat quality. In other words, ancient forest species such as Anemone nemorosa,
Oxalis acetosella or Hyacinthoides non-scripta in Europe hardly colonize recent
forest patches, not because they cannot establish but because diaspores hardly
disperse from source populations to recently created forest habitats, especially in
fragmented landscapes. In comparison, shade-tolerant species that are not
dispersal-limited (e.g. Urtica dioica, Galium aparine or Veronica hederifolia) can
colonize recent and ancient forests equally well. This colonization capacity has even
been quantified by an index ranging from minus 100 to 100 (reviewed in Flinn and
Vellend 2005; Hermy and Verheyen 2007). This example illustrates an important
step in the development of historical ecology as a science, since process-based
hypotheses are now experimentally tested.

Historical ecology also leads to hot debates in ecology. A remarkable example is


the controversy about what the primeval European forest used to look like. The
dominant idea has long been the closed-canopy forest hypothesis, mostly based on
the forest dynamics observed in nature reserves, once any significant human
pressure is removed, as well as on pollen records. This hypothesis has been
convincingly challenged by the savannah hypothesis introduced by Rackham (1998)
and further supported by a number of arguments reviewed in Vera (2000). The main
argument of the latter hypothesis is that many species of big herbivores used to
graze in forests (e.g. buffalos, aurochs, wild horses), and likely maintain the forested
areas quite open. Most of these species went extinct in the early Medieval times,
so that natural succession conducted to closed-canopy forests with a typically
shade-tolerant understory flora. Beyond the fact that these hypotheses are still
controversial (e.g. Birks 2005), both emerged as a result of historical ecology
approaches of forests (see also Chapter 13).

Furthermore, historical ecology tackles timely topics in ecology, for example the
impact of global climate changes on ecosystems. Progress in ecoinformatic has
rendered possible the use of large database compiling old ecological records. For
example, at least since the end of the 19th century, vegetation scientists have
accumulated millions of vegetation relevés worldwide. When these relevés can be
relocalized, it is possible to resurvey the same plots decades later (e.g. Verheyen
et al. 2017). The comparison between old and new relevés then makes it possible to
quantify vegetation changes and subsequently infer the drivers of these changes, by
using trait-based approaches (e.g. Closset-Kopp et al. 2019) or by computing
correlation with measured changes in environmental factors (e.g. Perring et al.
2016). Although inferring processes from resurvey studies is not trivial (see
Chapter 2), legacy studies led to important results, for example: the overriding
influence of canopy closure on atmospheric nitrogen deposits in explaining
vegetation eutrophication (Verheyen et al. 2012); the importance of land-use
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
6 Historical Ecology

legacies in explaining the response of forest biodiversity to global change (Perring


et al. 2016); and the crucial role of forest management in preserving forest
microclimate to buffer against global warming and conserve the forest understory
biodiversity (De Frenne et al. 2014; Bernhardt-Römermann et al. 2015; see also
Chapter 20). Long-term monitoring studies provide further support to the inferences
made from these retrospective analyses (see Chapter 3). Historical ecology has
evidenced the inertia of ecosystem in their response to environmental changes and
led to novel concepts that largely permeate ecology, for example the concepts of
extinction debt and colonization credit (Kuussaari et al. 2009; Jackson and Sax
2010; see also Chapter 21).

1.4. The way forward

If the roots of historical ecology are in Europe, its ramifications now reach all
continents. Several international working groups dedicated to historical ecology
emerged, within existing ecological societies such as the International Association
for Landscape Ecology (IALE; https://www.landscape-ecology.org/page-18083) and
the International Association for Vegetation Science (IAVS; https://www.iavs.org/
page/working-groups_historical-vegetation-ecology), and as standalone associations,
such as the American Society for Environmental History (https://aseh.org) and its
European counterpart (http://eseh.org). This international spread of historical
ecology is associated with a diversification of the studied systems. Temperate and
Mediterranean forests are still an important focus (see Chapters 3, 9, 10, 11, 12, 14,
15 and 20), but this book shows that historical ecology now concerns all types of
ecosystem and landscape: boreo-nemoral forests (Chapter 16), arctic (Chapter 17)
and tropical (Chapter 5) vegetation, temperate island (Chapter 18), subtropical
coastal dunes (Chapter 19), always with important insights for the understanding of
how past interactions between natural ecosystems and human societies have driven
current biodiversity patterns and ecosystem functioning.

The challenge for historical ecology since the beginning has been to understand
present-day ecosystems by putting them in their historical context. Moreover,
historical ecology becomes an applied discipline, for example, to identify reference
plant communities needed by restoration ecology practitioners (Swetnam et al. 1999;
Egan and Howell 2001; Foster 2002; Jackson and Hobbs 2009) and, more generally,
to use the past to predict the future (see Chapter 22). It is thus becoming a field
which informs stakeholders and managers. Landscapes and ecosystems cannot be
reduced to natural ecological systems disturbed by human activities. Instead, they
are dynamic socio-ecosystems in which human activities are both causes and
consequences of the observed patterns through time and across spatial scales. This is
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
A General Introduction to Historical Ecology 7

a timely task at the Anthropocene epoch, since understanding drivers of past


ecosystem changes may allow us to predict future changes in response to interacting
environmental and human drivers. Time has come to understand the mechanisms
behind these changes, in order to open the way for a functional historical ecology.
Historical ecology has the potential to become a “transdiscipline” dedicated to the
study of complex socio-ecosystem dynamics.

1.5. References

Balée, W. (2006). The research program of historical ecology. Annu. Rev. Anthropol., 35,
75–98.
Beck, A. (2011). Archaeological applications of multi/hyper-spectral data – Challenges and
potential. In Remote Sensing for Archaeological Heritage Management, Cowley, D. (ed.).
Archaeolingua, Budapest.
Beller, E., McClenachan, L., Trant, A., Sanderson, E.W., Rhemtulla, J., Guerrini, A.,
Grossinger, R., Higgs, E. (2017). Toward principles of historical ecology. Am. J. Bot.,
104, 645–648.
Bernhardt-Römermann, M., Baeten, L., Craven, D., De Frenne, P., Hédl, R., Lenoir, J., Bert,
D., Brunet, J., Chudomelová, M., Decocq, G. et al. (2015). Drivers of temporal changes in
temperate forest plant diversity vary across spatial scales. Glob. Change Biol., 21,
3726–3737.
Birks, H.J.B. (2005). Mind the gap: How open were European primeval forests? Tr. Ecol.
Evol., 20, 154–156.
Birks, H.J.B. and Birks, H.H. (2016). How have studies of ancient DNA from sediments
contributed to the reconstruction of Quaternary floras? New Phytol., 209, 499–506.
Challis, K., Kokalj, Z., Kincey, M., Moscrop, D., Howard, A.J. (2008). Airborne lidar and
historic environment records. Antiquity, 82, 1055–1064.
Closset-Kopp, D., Hattab, T., Decocq, G. (2019). Do drivers of forestry vehicles also drive
herb layer changes (1970–2015) in a temperate forest with contrasting habitat and
management conditions? J. Ecol., 107, 1439–1456.
De Frenne, P., Rodriguez-Sanchez, F., Coomes, D.A., Baeten, L., Verstraeten, G., Vellend,
M., Bernhardt-Römermann, M., Brown, C.D., Brunet, J., Cornelis, J. et al. (2014). Forest
canopy closure buffers plant community responses to global warming. Proc. Nat. Acad.
Sci. USA, 110, 18561–18565.
Decocq, G. (2004). Utilisation de la flore et de la végétation actuelles en prospection
archéologique. In Méthodes et initiations d’histoire et d’archéologie, Racinet, P. and
Schwerdroffer, J. (eds). Editions du Temps, Nantes.
Dupouey, J.L., Dambrine, E., Laffite, J.D., Moares, C. (2002). Irreversible impact of past land
use on forest soils and biodiversity. Ecology, 83, 2978–2984.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
8 Historical Ecology

Egan, D. and Howell, E.A. (eds) (2001). The Historical Ecology Handbook: A
Restorationist’s Guide to Reference Ecosystems. Island Press, Washington.
Feiss, T., Horen, H., Brasseur, B., Buridant, J., Gallet-Moron, E., Decocq, G. (2017).
Historical ecology of lowland forests: Does pedoanthracology support historical and
archaeological data? Quat. Inter., 457, 99–112.
Fitzpatrick, S.M. and Keegan, W.F. (2007). Human impacts and adaptations in the Caribbean
Islands: An historical ecology approach. Earth Env. Sci. Trans. Roy. Soc. Edinburgh, 98,
29–45.
Flinn, K.M. and Vellend, M. (2005). Recovery of forest plant communities in
post-agricultural landscapes. Front. Ecol. Env., 3, 243–250.
Foster, D.R. (2002). Conservation issues and approaches for dynamic cultural landscapes.
J. Biogeograph., 29, 1533–1535.
Hermy, M. and Verheyen, K. (2007). Legacies of the past in the present-day forest
biodiversity: A review of past land-use effects on forest plant species composition and
diversity. Ecol. Res., 22, 361–371.
Jackson, S.T. and Hobbs, R.J. (2009). Ecological restoration in the light of ecological history.
Science, 325, 567–569.
Jackson, S.T. and Sax, D.F. (2010). Balancing biodiversity in a changing environment:
Extinction debt, immigration credit and species turnover. Tr. Ecol. Evol., 25, 153–160.
Kuussaari, M., Bommarco, R., Heikkinen R.K., Helm, A., Krauss, J., Lindborg, R., Öckinger,
E., Pärtel, M., Pino, J., Rodà F. et al. (2009). Extinction debt: A challenge for biodiversity
conservation. Tr. Ecol. Evol., 24, 564–571.
Maury, A. (1850). Histoire des grandes forêts de la Gaule et de l’ancienne France.
A. Leleux, Paris.
Nash, R.F. (1967). Wilderness and the American Mind. Yale University Press, New Haven.
Perring, M.P., De Frenne, P., Baeten, L., Maes, S.L., Depauw, L., Blondeel, H., Caron, M.M.,
Verheyen, K. (2016). Global environmental change effects on ecosystems: The
importance of land-use legacies. Glob. Change Biol., 22, 1361–1371.
Peterken, G.F. (1981). Woodland Conservation and Management. Chapman & Hall, London.
Plue, J., Hermy, M., Verheyen, K., Thuillier, P., Saguez, R., Decocq, G. (2008). Persistent
changes in forest vegetation and seed bank 1,600 years after human occupation. Landsc.
Ecol., 23, 673–688.
Rackham, O. (1980). Ancient Woodland: Its History, Vegetation and Uses in England.
Edward Arnold, London.
Rackham, O. (1998). Implications of historical ecology for conservation. In Conservation
Science and Action, Sutherland, W.J. (ed.). Blackwell Science Oxford, Oxford.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
A General Introduction to Historical Ecology 9

Rawlence, N.J., Lowe, D.J., Wood, J.R., Young, J.M., Churchman, G.J., Huang, Y.T.,
Cooper, A. (2014). Using palaeoenvironmental DNA to reconstruct past environments:
Progress and prospects. J. Quatern. Sci., 29, 610–626.
Redman, C., Grove, M.J., Kuby, L. (2004). Integrating social science into the long term
ecological research (LTER) network: Social dimensions of ecological change and
ecological dimensions of social change. Ecosystems, 7, 161–171.
Russell, E.W.B. (1997). People and the Land through Time: Linking Ecology and History.
Yale University Press, New Haven.
Schoonmaker, P.K. and Foster, D.R. (1991). Some implications of paleoecology for
contemporary ecology. Bot. Rev., 57, 204–245.
Stisser, F.U. (1737). Forst- und Jagd-Historie der Teutschen. Ritter, Jena.
Swetnam, T.W., Allen, C.D., Betancourt, J.L. (1999). Applied historical ecology: Using the
past to manage for the future. Ecol. Applic., 9, 1189–1206.
Szabó, P. (2015). Historical ecology: Past, present and future. Biol. Rev., 90, 997–1014.
Vera, F.W.M. (2000). Grazing Ecology and Forest History. CABI Publishing, Wallingford.
Verheyen, K., Baeten, L., De Frenne, P., Bernhardt-Römermann, M., Brunet, J., Cornelis, J.,
Decocq, G., Dierschke, H., Eriksson, O., Hedl, R. et al. (2012). Driving factors behind the
eutrophication signal in understorey plant communities of deciduous temperate forests.
J. Ecol., 100, 352–365.
Verheyen, K., De Frenne, P., Baeten, L., Waller, D., Hédl, R., Perring, M., Brunet, J.,
Chudomelova, M., De Lombaerde, E., Depauw, L. et al. (2017). Combining community
resurvey data to advance global change research. BioScience, 67, 73–83.
Wilkinson, K. and Stevens, C. (2003). Environmental Archaeology. Approaches, Techniques
and Applications. Tempus Publishing, Stroud.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2

Historical Resurveys
Reveal Causes of Long-term
Ecological Change
Donald M. WALLER
University of Wisconsin, Madison, USA

2.1. Serious ecological changes are pervasive

Historical approaches have always informed ecology but have gained great
importance now as we struggle to understand the forces driving the many ecological
changes occurring around us. These changes are pervasive and accelerating, yet they
remain largely hidden as they are difficult to track. Documenting even simple
changes in diversity is impossible except where we have historical data. One of
Europe’s revered natural areas, the Bialowieza forest in Poland, lost 45% of its 133
vascular plant species between 1969 and 1992 (Kwiatkowska 1994). Two protected
virgin forests in Pennsylvania lost 59% and 80% of their plant species between 1929
and 1995 (Rooney and Dress 1997). Three state parks in northern Wisconsin, USA,
lost 50% of their plant diversity over the past 50 years (Rooney et al. 2004).
Middlesex Fells, a 400 ha park near Boston, USA, lost 37% of its 422 plant species
between 1894 and 1993 (Drayton and Primack 1996). Although losing species from
small particular sites may seem inconsequential, these studies suggest we are losing
species from protected temperate forests around the world.

What causes these losses? Are they particular to these sites or occurring in most
temperate forests? Are similar forces at work in these cases or is each different?
Although these results are dramatic, they would entirely escape notice except for our

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
12 Historical Ecology

having historical data at these sites. Knowing how these forces affect community
dynamics improves our ability to identify the particular forces that most threaten
biodiversity and is thus essential for wisely managing and restoring these
ecosystems.

2.2. Anthropogenic drivers of ecological change

Plant communities respond to many anthropogenic forces. Humans have driven


ecological changes for millennia via fire, herding ungulates and tilling fields.
However, the scope and rates of anthropogenic environmental change escalated
greatly through the 20th century. These forces continue to expand, ensuring that they
will increasingly affect ecosystems through the 21st century. Here, I focus on six
anthropogenic forces known to greatly affect plant communities. I cite examples
from long-term resurveys of sites in Wisconsin, USA, and forestREplot
(forestreplot.ugent.be/) resurveys in Europe to illustrate how these forces affect
temperate forests. The Wisconsin work capitalizes on high-quality baseline surveys
by J.T. Curtis (1959) and his students matched to subsequent resurveys of the same
sites 50+ years later (Waller et al. 2012). In temperate forests, most plant species
occur in the herb layer (Gilliam et al. 2016). Their shorter stature and life cycles
mean that these species respond faster to anthropogenic forces than forest trees.

I first discuss difficulties in studying long-term ecological change.


Understanding these help us to explain why good long-term studies are scarce. I then
review the different kinds of ecological change that exist and methods to
characterize these. These include natural and manipulative experiments,
metacommunity models and various statistical approaches. With this background, I
then review evidence linking ecological changes to shifts in disturbance regimes,
habitat fragmentation, climate change, nitrogen deposition and grazing by
herbivores. These concrete examples help us understand the strengths and
weaknesses of each approach. Finally, I summarize our current understanding of the
forces driving ecological change and hint at new approaches that might help us
advance further.

2.2.1. The missing baseline problem

Without adequate baseline data, we cannot detect long-term ecological change or


analyze how systems have changed (Magnuson 1990). Missing baselines also
prevent us from measuring species losses and biotic homogenization, shifts in
community composition and other long-term changes. Worse, the absence of data is
often taken to mean an absence of effects. In the few cases where we do have
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Resurveys Reveal Causes of Long-term Ecological Change 13

detailed baseline data, we often find serious biotic declines (see opening). It
behooves us to seek out historical data whenever possible, to archive them carefully
and encourage their wider use (Verheyen et al. 2017). Baseline data are often limited
to simple species lists as in the opening examples. Such lists identify the species
present but cannot inform us about shifts in relative abundance or community
structure. It is more useful to have quantitative data on species abundances at several
times (Figure 2.1). More extensive data enable more general and robust conclusions
about the kinds of changes occurring and how consistent these changes are. Having
data from multiple sites and species also presents challenges; however, in that
effects of local circumstances can be complex. Larger datasets provide more power
for increasingly sophisticated analyses, allowing us to better understand the patterns
and processes of ecological changes.

One species Many species

One Trend?
Multiple
site trends

Many Testable
Multiple
sites trend testable
trends
Time Time

Figure 2.1. Our ability to characterize ecological change and reliably infer causes
rests on having adequate baseline data from many species and sites and a suitable
interval between surveys. Such data are rare. For a color version of this figure, see
www.iste.co.uk/decocq/ecology.zip

2.2.2. Ecological communities are complex

Many forces drive long-term ecological change, often in combination,


complicating our efforts to infer them. These mechanisms operate at different times
or places, changing at various scales and rates (Bernhardt-Römermann et al. 2015).
Drivers also interact in complex ways, complicating our efforts to infer their relative
strengths and patterns of causality. Finally, it is difficult to distinguish systematic
patterns of change attributable to any given driver from stochastic changes driven by
shifts in local conditions and chance events. Having rich data from many sites
(Figure 2.1) gives us the statistical power we need to disentangle multiple drivers
while averaging over variable local conditions.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
14 Historical Ecology

2.3. Kinds of ecological change

Ecologists now use historical data to track many kinds of ecological change.
These include changes in species abundance and relative abundance; local
colonizations and extirpations; invasions by non-native species; changes in
community and habitat structure (e.g. plant density, cover and stature; changes in
local (α), among-site (β) and regional (γ) diversity; altered relationships with other
species (e.g. the incidence of herbivory and diseases); and shifts in higher-order
meta-community properties. At the community level, we track shifts in relative
abundance, plant functional traits, species–area relationships, etc. Studies usually
report just one or a few types of change, making it difficult to infer how these
various kinds of change relate to each other.

2.3.1. Natural community dynamics

To detect and analyze the effects of anthropogenic drivers of ecological change,


we must be able to separate those from natural processes affecting community
dynamics (as with stochastic noise). Classic ecological work has documented how
plant communities respond to variation in climate, soils and patterns of disturbance.
(e.g. Curtis 1959). Well-described baseline conditions help here, too. European
ecologists use Ellenberg indicator values from indicator species to infer
environmental conditions (Ellenberg 1991; Bartelheimer and Poschlod 2016).
Recurring natural disturbances including tree falls, fires and floods restart
succession, generating well-characterized changes in forest plant communities.

2.3.2. Anthropogenic drivers of ecological change

2.3.2.1. Shifts in disturbance regimes


Farm fields, managed forests, houses, roads and urban areas have displaced
prairies, savannas, wetlands and wild forests across much of temperate North
America, especially in areas favorable for agriculture. These shifts in land-use
greatly reduce natural disturbances like floods and fires, decimating populations of
species dependent on these disturbances. In contrast, species adapted to
human-disturbed environments thrive. This group includes many invasive species
which can themselves displace native species (see section 3.2.5). Species adapted to
historically dominant patterns of natural disturbance do poorly follow changes in the
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Resurveys Reveal Causes of Long-term Ecological Change 15

type, frequency, scale or intensity of historical disturbance regimes. This


undermines the ability of many species to thrive, leading to the pronounced declines
in diversity observed when burning declines in fire-dominated systems like prairies
(Alstad et al. 2016), savannas and jack-pine-dominated forests on sandy soils
(Li and Waller 2015). Floodplain communities are similarly disrupted when
flooding either declines or becomes more extreme (Johnson et al. 2016). Recurrent
logging similarly threatens many native species by failing to generate the same kinds
and amounts of gaps, tip-up mounds, soil structure and coarse woody debris as
natural death and tree-falls do in mature temperate forests (Alverson et al. 1994).

2.3.2.2. Habitat loss and fragmentation


GIS analyses of aerial imagery allow us to evaluate how landscape features
affect patterns of diversity. Forest patches in southern Wisconsin, like those in other
temperate areas, are now smaller and more isolated than they once were. This
reduces the abundance of species sensitive to edge, area or isolation effects leading
to many extirpations. Edge effects penetrate far into mature forests. The densities of
roads and houses within a 5 km radius cause native plant diversity to decline and
invasions of exotic species to increase (Rogers et al. 2008). As species that need
larger areas or less disturbed conditions disappear locally, the diversity of forest
patches declines. Over the latter half of the 20th century, this considerably reduced
the diversity of smaller forest patches in southern Wisconsin (Figure 2.2). This
strengthened the species–area relation matching how land-bridge islands lose
diversity after they become isolated, a tenant of island biogeography (MacArthur
and Wilson 1967). Smaller forest patches seem destined to continue losing species if
they remain small, isolated and subject to other stresses.

2.3.2.3. Atmospheric nitrogen deposition


Temperate ecosystems in proximity to industrialized urban areas or intensive
agriculture are often affected by atmospheric nitrogen (N) deposition. The wet and
dry deposition of oxidized forms of N and sulfur cause “acid rain”, leaching needed
cations and mobilizing toxic metals. This harms many species. Farmers apply
reduced forms of N (as ammonium compounds) as fertilizer, some of which escapes
as drift. Higher ambient levels of biologically available N favor plant species that
thrive under high N conditions and disfavor plants (like many Fabaceae) that either
fix N or perform well under low nutrient conditions (Verheyen et al. 2012). Such
“terrestrial eutrophication” commonly acts to simplify plant communities by
favoring fast-growing N-loving plants that can then overtop and outcompete more
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
16 Historical Ecology

nutrient-thrifty and shorter-statured plant species (Stevens et al. 2018). Across the
temperate U.S., higher N deposition acts to depress plant diversity particularly on
low nutrient soils (Simkin et al. 2016). Over half the 348 species evaluated are
vulnerable to N deposition including many rare habitat specialists and species with
low leaf N (Clark et al. 2019). These results confirm that N deposition has driven
widespread changes in many plant communities (Figure 2.3).

Figure 2.2. The number of species sustained in smaller forest fragments in southern
Wisconsin declined greatly through the latter 20th century. This strengthened the
species–area relationship. Data from Rogers et al. (2008)

2.3.2.4. Climate change


Rates of climate change are accelerating, resulting in sizable shifts in rainfall,
temperature and seasonality. Shifts in climate are now altering leaf and flowering
phenology, shifting elevational and latitudinal ranges, and increasing risks of local
extirpation (Parmesan and Yohe 2003). In Wisconsin, night and winter temperatures
have increased the most, with local increases in precipitation and more frequent
extreme storms (WICCI 2010). These shifts in climate have already shifted the
distributions of many species relative to the mid-20th century (Ash et al. 2016).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Resurveys Reveal Causes of Long-term Ecological Change 17

These rates of movement, however, are slower than those of climate change. This
creates a growing mismatch between species distributions and ideal climatic
conditions. This lag will likely to increase, threatening the ability of many species to
persist.

2.3.2.5. Invasions by non-native species


Many species have spread far beyond their original ranges to new regions either
by hitching rides with human commerce or via intentional introductions. Some
species continue to spread and have become invasive. The absence of coevolved
predators and pathogens in their introduced range may play a role in these
expansions, particularly when species reallocate resources from defense to
competitive ability and enhanced reproduction (Blossey and Notzold 1995).
Introduced plants compete for resources and alter the environment, shifting species
composition and potentially threatening native species (Callaway and Aschehoug
2000). Prominent invaders of southern Wisconsin forests include Alliaria petiolata,
Berberis thunbergii and Rhamnus cathartica. These species reduce plant diversity,
tree regeneration and the nesting success of birds. Introduced insects like the
Hemlock Woolly Adelgid and the Emerald Ash Borer are eliminating Tsuga
canadensis and several Fraxinus species from broad regions. Introduced diseases
like Chestnut Blight, Dutch Elm Disease and Beech Bark Disease have already
massively altered North American forests.

2.3.2.6. Hyper-abundant herbivores


Deer (cervids) became far more abundant across much of North America and parts
of Europe through the 20th century, with diverse and often dramatic cascading impacts
on many understory herbaceous and woody plant species as well as forest structure,
soil dynamics, and bird and mammal abundances (Alverson et al. 1988; Waller and
Alverson 1997; Côté et al. 2004). In northern Wisconsin, species known to resist or
tolerate herbivory by white-tailed deer (grasses, sedges and some ferns) increased
dramatically over the past 50 years. In contrast, most native forbs and species known
to be susceptible to herbivory (e.g. lilies and orchids) suffered large declines
(Balgooyen and Waller 1995). Geographic patterns confirm that deer drive long-term
shifts in community composition. Islands lacking deer and Indian reservations with
few or no deer still support high diversity including species highly palatable to deer
now rare elsewhere. Fenced exclosures confirm how deer radically change the cover,
diversity and composition of understory plants including the ability of several
late-successional species to effectively recruit saplings. Once deer deplete the
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
18 Historical Ecology

abundance of palatable species, the scarcity of seed sources and the tendency of deer
to seek out favored species means these species are slow to recover.

2.3.2.7. Drivers interact


Although each driver has substantial effects by itself, these drivers also interact.
For example, abundant deer tend to favor invasions by weedy plants like Alliaria
and Rhamnus by compacting soils, accelerating nutrient cycling, dispersing seeds
and preferentially consuming their native competitors (Dobson and Blossey 2015;
Shen et al. 2016). One aspect of widespread biotic homogenization in Europe is the
replacement of smaller-ranged plant species by those with larger ranges (Staude
et al. 2020). Interestingly, these larger-ranged species have higher leaf N and thus
probably benefit from N deposition. Additional key interactions surely exist among
the drivers reviewed here. Ignoring such interactions leads us to underestimate the
scale and complexity of the forces driving ecological change.

2.4. Understanding the forces driving ecological change

Scientists reduce their uncertainty by formulating hypotheses, conducting


experiments and analyzing the results from those experiments to test hypotheses.
Given the number and scale of human-related factors affecting ecological systems in
the 21st century, the scarcity of baseline data, time delays in these effects and how
complex these responses may be, we cannot conduct experiments to manipulate all
relevant factors. Instead, ecologists infer these forces using various direct and
indirect methods.

2.4.1. Natural experiments

Particular patterns of change often give us clues about the mechanisms driving
change. We compare sites that differ in local conditions or landscape context to see
how these affect the ecological changes observed. In cases where we have systematic
differences among regions or locales in environmental conditions, we analyze these
differences as “treatments” in a “natural experiment”, for example, by comparing
watersheds with or without dams (Johnson et al. 2016) or areas with and without
invasive earthworms. Comparing islands in Lake Superior with or without deer
confirmed their strong influence (Balgooyen and Waller 1995). Indian reservations
also provide natural experiments for assessing the effects of alternative paths of forest
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Resurveys Reveal Causes of Long-term Ecological Change 19

and wildlife management (Waller and Reo 2018). When many sites are analyzed
together that cover a broad range of conditions, we may gain enough power to perform
factorial natural experiments to test multiple drivers and their interactions. We can also
relate species’ functional traits to their within and among-site dynamics (Waller et al.
2017).

2.4.2. Metrics of change

By tracking multiple species or whole communities, we can tally local species


colonizations and extinctions or increases and decreases in plant size, density or
abundance. On the community level, we can track changes in structure and stature as
well as species (and trait) composition or overall diversity. Diversity has many
components and can be measured in many ways. Spatially, we distinguish among local
(α), among-site (β) and overall regional (γ) diversity. Beta diversity reflects
differences among biotic communities and thus provides an inverse measure of “biotic
homogenization” (McKinney and Lockwood 1999). These concepts and measures
have now been sorted into a coherent theory based on Hill numbers (Jost 2007), with
natural extensions to measure trait and phylogenetic diversity (Chao et al. 2019).

2.4.3. Can functional traits reveal drivers of change?

Plant functional traits have been hailed as a vehicle to “rebuild” community


ecology (McGill et al. 2006). Because plant functional traits reflect how plants
respond to environmental conditions (Rolhauser et al. 2021), we gain insights into
the forces driving shifts in community composition by examining how traits shift as
species rise or fall in abundance. Knowing how traits affect physiological and
ecological performance under various conditions allows us to predict how changed
conditions likely affect species with different trait profiles (Figure 2.3). For
example, systematic increases in nitrogen (N) deposition will tend to favor
faster-growing plants with higher leaf N and specific leaf area (fitting many invasive
plant species). Habitat fragmentation, in contrast, tends to favor species that disperse
well, for example, with small seeds or those that attach, or are eaten by, animals.
Deer feed selectively on taller plants with thin nutritious leaves, leaving in their
wake plants with better physical and chemical defenses. Tracking shifts in
community trait profiles allows us to test these predictions. Researchers now use
functional traits to track how communities respond to N-enrichment (Suding et al.
2005), climate change (Heilmeier 2019; Thomas et al. 2020), habitat fragmentation
(Vellend et al. 2006) and various kinds of disturbance (Mayfield et al. 2010).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
20 Historical Ecology

Figure 2.3. Factors that alter plant communities affect species differently depending
on their functional traits. This can generate diagnostic trait “signatures”, allowing us
to discriminate among drivers by tracking shifts in community trait profiles. For a color
version of this figure, see www.iste.co.uk/decocq/ecology.zip

2.4.4. Vectors of change – ordination

Since the days of Whittaker and Curtis in the 1950s, ecologists have used
multivariate methods to analyze plant community composition (Shipley 2021).
Multivariate approaches are also useful for analyzing how communities respond to
drivers of ecological change. For example, ordination allows us to visualize and
compare patterns of community change. Analyzing both baseline and resurvey data
together allows us to construct vectors reflecting how communities at each site have
shifted in composition (Figure 2.4).

We can then assess how these changes differ among sites, regions or community
types. Parallel vectors represent consistent responses. Converging vectors indicate
biotic homogenization. Forests in Wisconsin show somewhat parallel changes but
fragmented southern forests have clearly changed more. We can extend our use of
ordination by examining how environmental factors (e.g. soil or landscape
conditions) and plant traits covary (or “load”) with ordination axes. Environmental
factors that covary strongly with vectors of community (or trait) changes at
particular sites are candidate likely drivers of those changes. Shifts in plant traits
indicate how plant characteristics affect plant responses to drivers. These
multivariate approaches deserve wider application.
Figure 2.4. Vectors of community change. Left: NMDS ordinations showing how the composition of plant communities in the
fragmented deciduous forests of southern Wisconsin (solid points on left) and more continuous forests in northern Wisconsin
(empty circles on right) changed over the past half-century. Right: placing the origins of change vectors together reveals that
forests in both regions have shifted similarly, but that southern forests have changed more
Historical Resurveys Reveal Causes of Long-term Ecological Change
21

Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
22 Historical Ecology

2.5. Conclusion

Ecological changes continue to grow in number and extent. All the


environmental forces mentioned affect forest plant communities although their
relative importance varies across regions and community types (Table 2.1). Forces
like climate change are clearly growing in importance, while drivers like acid rain
may be declining. As drivers act slowly and cumulatively, they exert both direct and
indirect effects at multiple scales. This makes it hard to distinguish cause from effect
(“passengers” vs. “drivers”). Applying new methods to growing datasets will allow
us to better track the separate and combined impacts of different forces. It is
particularly important to understand cumulative impacts and how drivers interact. In
many areas, increases in clonal understory dominants like Circaea, Carex and
Parthenocissus create “recalcitrant understory layers” that suppress the growth of
tree seedlings and other species (Royo and Carson 2006). These layers often reflect
herbivory and altered disturbance regimes. We ignore these synergistic effects at our
peril. Surely, the effects of climate change will be compounded by habitat loss,
preventing the adaptive migration of many species and restructuring communities
(Radeloff et al. 2015).

The scarcity of historical baseline data remind us: today’s surveys provide the
bases for tomorrow’s resurveys. The more we can standardize our field methods,
which variables we track, and how we analyze our data, the more useful our results
become. Satellite and aerial imagery, data sharing groups like sPlot, and
standardized data from NEON (www.neonscience.org/) and the Forest Inventory
and Analysis program (www.fia.fs.fed.us/) give us “big data” sets for studying
ecological change. Synthesis centers like iDiv in Germany (www.idiv.de) are also
enhancing our ability to compile, analyze and interpret ecological data by integrating
vegetation plot surveys with climate, range and functional trait data (e.g. TRY,
www.try-db.org). These efforts are yielding new insights of great value as we strive
to understand, and adapt to, the forces driving ecological change.

2.6. References

Alstad, A.O., Damschen, E.I., Givnish, T.J., Harrington, J.A., Leach, M.K., Rogers, D.A.,
Waller, D.M. (2016). The pace of plant community change is accelerating in remnant
prairies. Sci. Adv., 2, e1500975.
Alverson, W.S., Waller, D.M., Solheim, S.L. (1988). Forests too deer: Edge effects in
northern Wisconsin. Conserv. Biol., 2, 348–358.
Alverson, W.S., Kuhlmann, W., Waller, D.M. (1994). Wild Forests: Conservation Biology
and Public Policy. Island Press, Washington, DC.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Resurveys Reveal Causes of Long-term Ecological Change 23

Ash, J.D., Givnish, T.J., Waller, D.M. (2017). Tracking lags in historical plant species’ shifts
in relation to regional climate change. Glob. Chang. Biolo., 23, 1305–1315.
Balgooyen, C.P. and Waller, D.M. (1995). The use of Clintonia borealis and other indicators
to gauge impacts of white-tailed deer on plant communities in northern Wisconsin, USA.
Nat. Areas J., 15, 308–318.
Bartelheimer, M. and Poschlod, P. (2016). Functional characterizations of Ellenberg indicator
values – A review on ecophysiological determinants. Funct. Ecol., 30, 506–516.
Bernhardt-Römermann, M., Baeten, L., Craven, D., De Frenne, P., Hédl, R., Lenoir, J.,
Bert, D., Brunet, J., Chudomelová, M., Decocq, G. et al. (2015). Drivers of temporal
changes in temperate forest plant diversity vary across spatial scales. Glob. Chang. Biol.,
21, 3726–3737.
Blossey, B. and Notzold, R. (1995). Evolution of increased competitive ability in invasive
nonindigenous plants: A hypothesis. J. Ecol., 83, 887–889.
Callaway, R.M. and Aschehoug, E.T. (2000). Invasive plants versus their new and old
neighbors: A mechanism for exotic invasion. Science, 290, 521–523.
Chao, A., Chiu, C.H., Villéger, S., Sun, I.F., Thorn, S., Lin, Y.C., Chiang, J.M.,
Sherwin, W.B. (2019). An attribute-diversity approach to functional diversity, functional
beta diversity, and related (dis)similarity measures. Ecol. Monogr., 89, e01343.
Clark, C.M., Simkin, S.M., Allen, E.B., Bowman, W.D., Belnap, J., Brooks, M.L.,
Collins, S.L., Geiser, L.H., Gilliam, F.S., Jovan, S.E. et al. (2019). Potential vulnerability
of 348 herbaceous species to atmospheric deposition of nitrogen and sulfur in the United
States. Nat. Plants, 5, 697–705.
Côté, S.D., Rooney, T.P., Tremblay, J.-P., Dussault, C., Waller, D.M. (2004). Ecological
impacts of deer overabundance. Annu. Rev. Ecol. Evol. Syst., 35, 113–147.
Curtis, J.T. (1959). The Vegetation of Wisconsin. University of Wisconsin Press, Madison.
Dobson, A. and Blossey, B. (2015). Earthworm invasion, white-tailed deer and seedling
establishment in deciduous forests of north-eastern North America. J. Ecol., 103,
153–164.
Drayton, B. and Primack, R.B. (1996). Plant species lost in an isolated conservation area in
metropolitan Boston from 1894 to 1993. Conserv. Biol., 10, 30–39.
Gilliam, F.S., Welch, N.T., Phillips, A.H., Billmyer, J.H., Peterjohn, W.T., Fowler, Z.K.,
Walter, C.A., Burnham, M.B., May, J.D., Adams, M.B. et al. (2016). Year response of the
herbaceous layer of a temperate hardwood forest to elevated nitrogen deposition.
Ecosphere, 7, 1–16.
Heilmeier, H. (2019). Functional traits explaining plant responses to past and future climate
changes. Flora, 254, 1–11.
Johnson, S.E., Amatangelo, K.L., Townsend, P.A., Waller, D.M. (2016). Large, connected
floodplain forests prone to flooding best sustain plant diversity. Ecology, 97, 3019–3030.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
24 Historical Ecology

Jost, L. (2007). Partitioning diversity into alpha and beta components. Ecology, 88,
2427–2439.
Kwiatkowska, A.J. (1994). Changes in the species richness, spatial pattern and species
frequency associated with the decline of oak forest. Vegetatio, 112, 171–180.
Li, D. and Waller, D.M. (2015). Drivers of observed biotic homogenization in pine barrens of
central Wisconsin. Ecology, 96, 1030–1041.
MacArthur, E.O. and Wilson, R.H. (1967). The Theory of Island Biogeography. Princeton
University Press, Princeton.
Magnuson, J.J. (1990). Long-term ecological research and the invisible present: Uncovering
the processes hidden because they occur slowly or because effects lag years behind
causes. Bioscience, 40, 495–501.
Mayfield, M.M., Bonser, S.P., Morgan, J.W., Aubin, I., McNamara, S., Vesk, P.A. (2010).
What does species richness tell us about functional trait diversity? Predictions and
evidence for responses of species and functional trait diversity to land-use change. Glob.
Ecol. Biogeogr., 19, 423–431.
McGill, B.J., Enquist, B.J., Weiher, E., Westoby, M. (2006). Rebuilding community ecology
from functional traits. TREE, 21, 178–185.
McKinney, M.L. and Lockwood, J.L. (1999). Biotic homogenization: A few winners
replacing many losers in the next mass extinction. Trends Ecol. Evol., 14, 450–453.
Parmesan, C. and Yohe, G. (2003). A globally coherent fingerprint of climate change impacts
across natural systems. Nature, 421, 37–42.
Radeloff, V.C., Williams, J.W., Bateman, B.L., Burke, K.D., Carter, S.K., Childress, E.S.,
Cromwell, K.J., Gratton, C., Hasley, A.O., Kraemer, B.M. et al. (2015). The rise of
novelty in ecosystems. Ecol. Appl., 25, 2051–2068.
Rogers, D.A., Rooney, T.P., Olson, D., Waller, D.M. (2008). Shifts in southern Wisconsin
forest canopy and understory richness, composition, and heterogentity. Ecology, 89,
2482–2492.
Rolhauser, A.G., Waller, D.M., Tucker, C.M. (2021). Complex trait‒environment
relationships underlie the structure of forest plant communities. J. Ecol., 109, 3794–3806.
Rooney, T.P. and Dress, W.J. (1997). Patterns of plant diversity in overbrowsed primary and
mature secondary hemlock northern hardwood forest stands. J. Torrey Bot. Soc., 124,
43–51.
Rooney, T.P., Wiegmann, S.M., Rogers, D.A., Waller, D.M. (2004). Biotic impoverishment
and homogenization in unfragmented forest understory communities. Conserv. Biol., 18,
787–798.
Royo, A.A. and Carson, W.P. (2006). On the formation of dense understory layers in forests
worldwide: Consequences and implications for forest dynamics, biodiversity, and
succession. Can. J. For. Res., 36, 1345–1362.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Resurveys Reveal Causes of Long-term Ecological Change 25

Shen, X., Bourg, N.A., McShea, W.J., Turner, B.L., Rooney, T., Dress, W., Rooney, T.,
Wiegmann, S., Rogers, D.A., Waller, D.M. et al. (2016). Long-term effects of
white-tailed deer exclusion on the invasion of exotic plants: A case study in a Mid-
Atlantic temperate forest. PLoS One, 11, e0151825.
Shipley, B. (2021). Ordination Methods for Biologists: A Non-Mathematical Introduction
Using R. BS Publishing, Sherbrooke.
Simkin, S.M., Allen, E.B., Bowman, W.D., Clark, C.M., Belnap, J., Brooks, M.L., Cade, B.S.,
Collins, S.L., Geiser, L.H., Gilliam, F.S. et al. (2016). Conditional vulnerability of plant
diversity to atmospheric nitrogen deposition across the USA. Proc. Natl. Acad. Sci. USA,
113, 4086–4091.
Staude, I.R., Waller, D.M., Bernhardt-Römermann, M., Bjorkman, A.D., Brunet, J.,
De Frenne, P., Hédl, R., Jandt, U., Lenoir, J., Máliš, F. et al. (2020). Replacements of
small- by large-ranged species scale up to diversity loss in Europe’s temperate forest
biome. Nat. Ecol. Evol., 4, 802–808.
Stevens, C.J., David, T.I., Storkey, J. (2018). Atmospheric nitrogen deposition in terrestrial
ecosystems: Its impact on plant communities and consequences across trophic levels.
Funct. Ecol., 32, 1757–1769.
Suding, K.N., Collins, S.L., Gough, L., Clark, C., Cleland, E.E., Gross, K.L., Milchunas, D.G.,
Pennings. S. (2005). Functional- and abundance-based mechanisms explain diversity loss
due to N fertilization. Proc. Natl. Acad. Sci. USA, 102, 4387–4392.
Thomas, H.J.D., Bjorkman, A.D., Myers-Smith, I.H., Elmendorf, S.C., Kattge, J., Diaz, S.,
Vellend, M., Blok, D., Cornelissen, J.H.C., Forbes, B.C. et al. (2020). Global plant trait
relationships extend to the climatic extremes of the tundra biome. Nat. Commun., 111(11),
1–12.
Vellend, M., Verheyen, K., Jacquemyn, M.H., Kold, A., Peterken, G., Hermy, M. (2006).
Extinction debt of forest plants persists for more than a century following habitat
fragmentation. Ecology, 87, 542–548.
Verheyen, K., Baeten, L., De Frenne, P., Bernhardt-Römermann, M., Brunet, J., Cornelis, J.,
Decocq, G., Dierschke, H., Eriksson, O., Hédl, R. et al. (2012). Driving factors behind the
eutrophication signal in understorey plant communities of deciduous temperate forests.
J. Ecol., 100, 352–365.
Verheyen, K., De Frenne, P., Baeten, L., Waller, D.M., Hédl, R., Perring, M.P., Blondeel, H.,
Brunet, J., Chudomelová, M., Decocq, G. et al. (2017). Combining biodiversity resurveys
across regions to advance global change research. Bioscience, 67, 73–83.
Waller, D.M. and Alverson, W.S. (1997). The white-tailed deer: A keystone herbivore. Wildl.
Soc. Bull., 25, 217–226.
Waller, D.M. and Reo, N.J. (2018). First stewards: Ecological outcomes of forest and wildlife
stewardship by indigenous peoples of Wisconsin, USA. Ecol. & Soc., 23 [Online].
Available at: https://doi.org/10.5751/ES-09865-230145.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
26 Historical Ecology

Waller, D.M., Amatangelo, K., Johnson, S., Rogers, D.A. (2012). Wisconsin vegetation
database – Plant community survey and resurvey data from the Wisconsin plant ecology
laboratory. Biodivers. Ecol., 4, 255–264.
Waller, D.M., Mudrak, E.L., Rogers, D.A. (2017). Do metacommunity mass effects predict
changes in species incidence and abundance? Ecography, 41, 11–23.
WICCI (2010). Wisconsin initiative on climate change impacts [Online]. Available at:
http://www.wicci.wisc.edu/.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
3

Getting the Right Answer


Can Take a While: Long-term
Ecological Field Studies
as Historical Ecology
Frank S. GILLIAM
University of West Florida, Pensacola,USA

3.1. Introduction

Implicit in any discussion of the essential importance of historical ecology as an


ecological discipline is the dimension of temporal scale (Szabó and Hédl 2011).
Indeed, among the numerous challenges in ecological studies is that of addressing
both temporal and spatial scales, which Levin (1992) identified as a central problem
in ecology, one that unifies such otherwise-disparate subdisciplines as population
ecology and ecosystem ecology, as well as basic and applied ecology. For example,
essential ecosystem drivers and related distribution of vegetation operate on spatial
scales from just over 0 to >10,000 km and temporal scales from hours to millennia.
Ecosystem drivers would include weather, climate variability/climate change, fire
and land-use change, whereas vegetation ranges from local patches to global biomes
(Figure 3.1, top). It is not coincidental that the spatial/temporal scales of historical
ecology parallel those of both the variety of ecological subdisciplines and the study
of environmental drivers and vegetation (Figure 3.1, bottom).

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
28 Historical Ecology

Figure 3.1. Spatial and temporal scales of (top) essential ecosystem drivers
(weather, climate variability and climate change, fire and land-use change) and
related distribution of vegetation. Figure reprinted from Gilliam (2016), used with
permission, and (bottom) subdisciplines of the study of ecology and of the study of
history. Figure reprinted from Szabó and Hédl (2011) with permission. For a color
version of this figure, see www.iste.co.uk/decocq/ecology.zip
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Long-term Ecological Field Studies as Historical Ecology 29

It is imperative to study ecological systems at the appropriate temporal scale, with


most field-based endeavors studying phenomena that operate under longer periods.
Such studies occasionally yield new questions to address when novel ecological
patterns and processes emerge through their extended duration. Perhaps, the best
example of this is the Park Grass Experiment, currently the longest ongoing field
experiment in the world (Richardson 1938). It was initiated in 1856 at Rothamsted in
Hertfordshire, England, by John B. Lawes and Joseph H. Gilbert, who wanted to
address a very pragmatic concern for English dairy farmers – how to improve hay
yield via the addition of a variety of fertilizers (Silvertown et al. 2006). Although it
took relatively little time to answer that basic question, maintaining the experiment
over a far longer period allowed researchers to address fundamental ecological
processes never considered by Lawes and Gilbert, including effects of plant biomass
and excess nitrogen on plant diversity, trophic dynamics of an herb-dominated
ecosystem and genetic drift in plant populations (Silvertown et al. 2006). In addition,
such extended research eventually led to an awareness of the importance of long-term
studies in general (Likens 1989), including whole research programs, such as the
Long-Term Ecological Research and Long-Term Research in Environmental
Biology programs of the National Science Foundation in the United States.

The purpose of this chapter is to examine the importance of maintaining


long-term studies to generate and test hypotheses regarding the ecology of terrestrial
ecosystems, in this case, a hardwood forest ecosystem of the central Appalachian
region of the United States, namely, the Fernow Experimental Forest (FEF), West
Virginia, U.S.A. As it is not uncommon for papers to be published based on three to
five years of field data, of particular interest is how extending field sampling to over
nearly 30 years has, in many cases, led to conclusions quite different from those
based on shorter time periods. In addition to considerable work carried out by
scientists at the USDA Forest Service Timber and Watershed Laboratory, Parsons,
West Virginia, where facilities associated with FEF are located, FEF has been the
site of numerous cooperative studies by scientists from colleges and universities
primarily in the eastern U.S. Rather than attempt to review all such studies, this
chapter covers primarily only work carried out by the Weeds and Dirt Laboratory at
Marshall University, Huntington, West Virginia.

3.2. Fernow Experimental Forest

3.2.1. Background

Forest ecology at FEF has a long and productive research history (Adams et al.
2006). Initial work was on silvicultural practices (Smith and Miller 1987), with the
collection of hydrochemical data for the long-term reference watershed (WS4)
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
30 Historical Ecology

beginning in 1960. A specific focus on the effects of acid deposition was initiated
via the Watershed Acidification Study (WAS), which began as a now-terminated
pilot study in 1987 on a watershed adjacent to FEF. In 1989, it was established on
FEF watersheds and is currently ongoing, although treatments ceased in 2019. A
distinctive feature of the WAS is that has involved a whole-watershed application of
simulated acidic deposition – as (NH4)2SO4 – via three aerial additions of year,
representing total N addition of 35 kg N ha-1 yr-1. This has been applied in powder
form by airplane (Adams et al. 2006).

3.2.2. Site description

Fernow occupies ~1,900 ha of the Allegheny Mountain section of the


unglaciated Allegheny Plateau near Parsons, West Virginia (39o03’16”N,
79o41’0”W). Precipitation at FEF averages approximately 1,430 mm yr-1, generally
higher during the growing season and increasing with elevation. As of the initiation
of the WAS, concentrations of acidic species in wet deposition (H+, SO4= and NO3-)
were among the highest in North America. Watershed soils are coarse-textured
Inceptisols (loamy-skeletal, mixed mesic Typic Dystrochrept) of the Berks and
Calvin series sandy loams derived from sandstone (Adams et al. 2006).

Dominant tree species on FEF watersheds vary with stand age.


Early-successional species such as black birch (Betula lenta L.), black cherry
(Prunus serotina Ehrh.) and yellow-poplar (Liriodendron tulipifera L.) are dominant
in young stands, whereas late-successional species, such as sugar maple (Acer
saccharum Marshall) and northern red oak (Q. rubra L.), are dominant in mature
stands. Dominant herbaceous layer species at the beginning of the study varied less
with stand age, including stinging nettle (Laportea canadensis (L.) Wedd.), violets
(Viola spp.) and several ferns (Gilliam et al. 2006).

3.2.3. Field design

Three watersheds are used for ongoing research as part of the WAS (Figure 3.2).
WS4 supports >100-year-old even-aged stands, serving as the long-time reference
watershed at FEF. WS7 supports an approximately 40-year-old even-age stand,
whereas WS3 supports an approximately 40-year-old even-age stand and serves as
the treatment watershed, whereas WS4 and WS7 are the controls. WS3 has received
three aerial applications of (NH4)2SO4 yr-1, beginning in 1989. March (or sometimes
April) and November applications represent approximately 7.1 kg N ha-1; July
applications are approximately 21.2 kg N ha-1. The total amount of N deposited on
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Long-term Ecological Field Studies as Historical Ecology 31

WS3 (application plus atmospheric deposition) is approximately 54 kg N ha-1 yr-1, or


about three times ambient inputs (Adams et al. 2006).

Figure 3.2. Experimental watersheds of the Fernow Experimental


Forest, West Virginia. WS3, WS4 and WS7 are used as
part of the ongoing Watershed Acidification Study

In 1991, 15 permanent vegetation plots were established in each of four watersheds


of contrasting stand age at FEF, with a particular focus on interactions between the
forest canopy and the herbaceous stratum, as well as plant–soil interactions, and how
these interactions might vary with forest succession following disturbance. A sub-set
of seven of these original plots in each of three watersheds – WS3, WS4 and
WS7 – were selected in 1993 for monthly in situ incubation of soil to examine spatial
and temporal patterns of net N mineralization and nitrification. This represented the
first work at FEF to specifically address a phenomenon that until that time had been
described more in Europe and experimental studies in the northeastern United States,
N saturation, which arises from atmospheric inputs of N exceeding biological N
demand (Aber et al. 1998). This typically results in the loss of NO3- in streams and is
commonly accompanied by the loss of nutrients (e.g. Ca++ and Mg++) (the mobile
anion effect) that are essential to plant growth and forest health. Some earlier studies
published on N saturation in the US (e.g. Stoddard 1994; Gilliam et al. 1996; Peterjohn
et al. 1996) identified watersheds at FEF to be among the more N-saturated sites in
North America. Symptoms included increased mobility and leaching of Ca++ and Mg++
as being linked to increased N deposition, along with enhanced nitrification and
movement of NO3- and decreases in growth rates of dominant tree species (Peterjohn
et al. 1996; Christ et al. 2002; May et al. 2005; Adams et al. 2006). Other problems
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
32 Historical Ecology

associated with N saturation include the increased production of the greenhouse gas,
N2O (Peterjohn et al. 1998). Further work at FEF has also suggested that N saturation
has led to phosphorous limitation in several watersheds (Gress et al. 2007).

3.3. Long-term studies at Fernow Experimental Forest, West Virginia

There are several examples from FEF wherein initial conclusions made after a
shorter period of investigation were contradicted by extending sampling over greater
time periods. In general, the original conclusion was a lack of effect of experimental
additions on treatment WS3, based on three to five years of treatment, often with a
focus on the N component of these additions. It should be noted, however, that the
numerous studies at other forest sites had, at that time, shown significant treatment
effects after a similar time period, especially regarding the response of the herb layer
(Hurd et al. 1998; Lu et al. 2010; Chapman et al. 2016), so the published lack of
treatment effects at FEF was notable in themselves. It should also be noted that some
of these studies focus on the experimental treatment as simulated acid deposition,
whereas others specifically focused on the N component of the treatment.

3.3.1. Effects of acidification on soil fertility and herb layer cover and
foliar nutrients

Gilliam et al. (1994) examined the effects of experimental acidification on soil


fertility and several characteristics of the herb layer using all three watersheds of the
WAS following a three-year treatment period for WS3. They found essentially no
significant differences in any of the measured soil variables – pH, organic matter,
cation exchange capacity, extractable nutrients – among watersheds, as well as no
differences in herb cover, biomass and foliar nutrients. This led to the conclusion
that three years of whole-watershed acidification had minimal, if any, effects on
nutrient availability – a surprising result at the time.

As a follow-up to Gilliam et al. (1994), in July 2015 – 24 years after the original
sampling in 1991 – Gilliam et al. (2020) resampled mineral soil from the 7 of the
original 15 plots in each of the three WAS watersheds. Although the focus was on
temporal shifts in effects of experimental acidification on soil fertility, they also
reported on extensive stream chemistry data, including concentrations of H+, NO3-,
base cations and electrical conductivity (Figure 3.3). Gilliam et al. (2020) yielded
sharply contrasting results from Gilliam et al. (1994), wherein virtually all measured
variables exhibited significant treatment effects. In general, pH was lower,
exchangeable acidity and aluminum were higher, nutrient (base) cations were lower,
and extractable N was higher following a quarter century of experimental acidification.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Long-term Ecological Field Studies as Historical Ecology 33

Figure 3.3. Volume-weighted mean monthly concentrations of major ions in stream


water and electrical conductivity from 1989 to 2014 for three small watersheds at the
Fernow Experimental Forest in West Virginia. Trend lines are 24-mo running means.
Vertical bars indicate when fertilizer applications began. Figure reprinted from
Gilliam et al. (2020) with permission. For a color version of this figure, see
www.iste.co.uk/decocq/ecology.zip
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
34 Historical Ecology

3.3.2. Effects of N addition on soil N dynamics

In 1993, the same subset of seven plots of the original 15 were identified for in
situ (“buried bag”) incubations for measurements of net N mineralization and
nitrification in surface mineral soil in each of the three WAS watersheds. Briefly,
this method involved sampling of mineral soil to a 5-cm depth at five locations in
each plot and combining/mixing soil to yield a single composite sample, which was
placed in two sterile polyethylene bags, one of which was buried at a 5-cm depth
and the other was taken to the lab for immediate extraction and analysis for NH4+
and NO3-. This was repeated monthly during the growing season on all 21 plots from
1993 to 1995, with each collection comprised of two bags: the buried bag from the
previous month and the bag from the current month. Differences in extractable N
between a non-buried bag and its buried equivalent were used to determine rates of
net N mineralization and nitrification.

Gilliam et al. (1996) reported no effects of experimental N additions on N


mineralization and nitrification for the first year of study, but found significant
seasonal variation. Gilliam et al. (2001) extended the sample period for the
three-year period of 1993–1995, finding significant, but transient, monthly effects
on net nitrification, wherein rates were significantly higher on WS3 for months that
immediately followed N additions. They also took advantage of the extended
sampling period to use multiple regression with ambient environmental variables,
concluding that net nitrification was directly influenced by mean ambient
temperature and soil moisture.

Following subsequent sampling in 2003 and 2005, the yearly sampling on these
plots was initiated in 2007 with funding from the NSF Long Term Research in
Environmental Biology (NSF LTREB) program to West Virginia University
(William T. Peterjohn, principal investigator). This work is currently ongoing.
Gilliam et al. (2018) reported on numerous aspects of N dynamics from the 21-year
period from 1993 to 2014, demonstrating that there had been a transient effect of N
additions on the annual mean net nitrification in 1995, after which all watersheds
exhibited similar declines to 2005 and are currently increasing. They found that the
best predictor of net nitrification was a metric of ambient temperature – number of
degree days <19°C – a significant pattern did not emerge until after the longer
period of study (Figure 3.4).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Long-term Ecological Field Studies as Historical Ecology 35

3.3.3. Effects of N addition on herb layer composition and diversity

Figure 3.4. Mean annual growing season net N mineralization (A) and net nitrification
(B) in mineral soil for study watersheds at Fernow Experimental Forest, West
Virginia, 1993–2014. Annual degree days < 19°C are shown. Fitted curves are
fifth-order polynomials for all variables to visually characterize temporal trends.
Degree days: closed circles/solid curve, r2 = 0.82; WS3: open circles/fine dashes,
r2 = 0.94 and 0.98 for net N mineralization and net nitrification, respectively; WS4:
open squares/intermediate dashes, r2 = 0.89 and 0.89; WS7: open triangles/wide
dashes, r2 = 0.96 and 0.95. Figure reprinted from Gilliam et al. (2018) with permission
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
36 Historical Ecology

Figure 3.5. Detrended correspondence analysis (DCA) for herbaceous layer


communities from 1991 to 2014 on reference WS4 (open symbols/dashed line)
and N-treated WS3 (solid symbols/solid line). Values shown are centroids
(two-dimensional means) for each year per watershed. Figure reprinted from Gilliam
et al. (2016) with permission

The initial conclusions regarding the effects of experimental acidification


(Gilliam et al. 1994) were based on a single year of sampling across all watersheds
(1991). Herb layer sampling was continued on WS3 and WS4 in 1992 and 1994,
representing a treatment period of three to six years on WS3, a time period during
which other studies had reported significant responses to experimentally added N
(e.g. Hurd et al. 1998; Lu et al. 2010). Thus, it was surprising that, in synthesizing
all data through 1994, Gilliam et al. (2006) reported no effects of added N on herb
cover, species richness, evenness and diversity at FEF. In addition, there was
minimal variation in species composition, both between treatment and reference
watershed and through time, from 1991 to 1994. They concluded that the lack of
treatment effect was from the high ambient N deposition at FEF, such that the
amount of added N (35 kg N/ha/yr) was a smaller percentage relative to ambient
deposition compared to other studies.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Long-term Ecological Field Studies as Historical Ecology 37

Herb layer sampling was repeated on a subset of 7 of the original 15 plots on WS3
and WS4 in 2004. Again, as a result of the NSF LTREB funding, annual sampling was
initiated in 2009 and is currently ongoing on WS3 and WS4 only. Gilliam et al. (2016)
reported results up to 2014, a total of 10 yr of data taken over a 23-year period. They
found significant and notable effects of excess N on all aspects of the herb layer
community, including increases in cover and decreases in species richness, evenness
and diversity (Figure 3.5). There were also profound changes in composition, with
cover of Rubus allegheniensis – once a minor component on all watersheds – increasing
>10-fold on N-treated WS3. The response of R. allegheniensis to N additions to WS3
has been confirmed by other studies at FEF (Walter et al. 2016, 2017).

3.3.4. The N homogeneity hypothesis

One of the more salient outcomes of long-term research at FEF was the
development – and eventual testing – of a hypothesis that addresses the effects of
excess N on forest herb communities, the N homogeneity hypothesis (Gilliam 2006).
The hypothesis predicts the loss of herb layer diversity from anthropogenically
increased N deposition via the following: (1) herb cover will initially increase (i.e. a
fertilizer effect), (2) N-mediated increases in herb cover simultaneously enhances
the growth of nitrophilous species and elimination of N-efficient species, (3) both
species richness and evenness will decline (i.e. from increased dominance of only
few nitrophilic species), contributing to declines in diversity, and (4) excess N will
increase the spatial homogeneity of N availability, contributing to increased sample
homogeneity of the herb layer (Gilliam et al. 2016). A key facet of the mechanism
behind this response is the importance of spatial heterogeneity of soil resources in
maintaining high diversity in plant communities (Hutchings et al. 2003).

Initial observations for the hypothesis began as early as Gilliam et al. (2001),
who reported lower spatial variation (measured as coefficients of variation of
watershed-scale means) in net nitrification and extractable NO3- pools in mineral soil
of treatment WS3 relative reference watersheds WS7 and WS4. It was originally
articulated as a predictive hypothesis in a review of responses of the herb layer of
forests of North America and Europe to excess N deposition (Gilliam 2006).
Although this hypothesis has received wide support in the literature (Cholewińska et
al. 2020), it was only able to be tested at FEF by extending the original sampling to
the current long-term period, including a more extensive one-time spatial sampling
of 100 plots in each of treatment WS3 and reference WS7 in 2011 (Figure 3.6). As
reported in Gilliam et al. (2016), the spatial homogeneity of net nitrification was
quite high on WS3, relative to WS7. Using the ongoing sampling from permanent
plots on all three watersheds, Gilliam et al. (2018) demonstrated N-mediated
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
38 Historical Ecology

increases in spatial homogeneity of net nitrification over time on WS3 up to a


quarter century following the initiation of treatments.

Figure 3.6. Spatial patterns of the potential net nitrification measured in 2011 using
surface soils (top 5 cm of mineral soil) collected from the reference (WS7) and
N‐treated (WS3) watersheds at the Fernow Experimental Forest, West Virginia.
Figure reprinted from Gilliam et al. (2016) with permission. For a color version of this
figure, see www.iste.co.uk/decocq/ecology.zip
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
39
Long-term Ecological Field Studies as Historical Ecology

reprinted from Gilliam et al. (2019) with permission. For a color version
Figure 3.7. Total deposition of NOx for 2000 (a) and 2014 (b). Figure

of this figure, see www.iste.co.uk/decocq/ecology.zip


Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
40 Historical Ecology

3.3.5. A look at the future: declines in the atmospheric deposition of N

During the decade preceding the initiation of the WAS, the wet deposition of N
at FEF averaged 7.3 kg N/ha/yr, whereas mean deposition during the most recent
decade has been 4.2 kg N/ha/yr, a decline of >40%. This decline has been reported
throughout North America (Du 2016; Lloret and Valiela 2016; Groffman et al. 2018;
Gilliam et al. 2019) and Europe (Schmitz et al. 2019), the result of regulatory
legislation in many parts of the world, such as the Clean Air Act in the U.S.
(Du et al. 2019). Because these global clean-air policies primarily target emissions
of N associated with energy combustion, declines in atmospheric N deposition are
far more prominent for NO3- than for NH4+ (Figure 3.7).

Such a trend does not call into question the relevance of long-term studies at FEF
or similar studies elsewhere. Rather, by maintaining such studies following cessation
of the N treatment on WS3, this work will become uniquely positioned to test
hypotheses regarding the legacies of chronic N deposition and N saturation
(Groffman et al. 2018; Gilliam et al. 2019; Schmitz et al. 2019). Groffman et al.
(2018) have suggested that impacted forests will undergo N oligotrophication as
increased carbon (C) flow from the atmosphere stimulates microbial N mobilization
in forest soil, decreasing available N for plants. Gilliam et al. (2019) proposed a
hysteretic model for decreased N future, wherein there will be widely varying time
lags in the recovery of soil acidification, plant biodiversity, soil microbial
communities, C and N cycling, and surface water chemistry towards pre-N impact
conditions. They concluded that, although there is a great deal of uncertainty
regarding forest response to future environmental change, it is quite certain that
ambient CO2 and temperature following any return to pre-impact status will not be
the same as they were prior to past increases in N deposition.

3.4. Conclusion

Historical ecology is inexorably linked with temporal scale, as it focuses on past


ecosystems and regards humans as one of numerous factors influencing these
systems (Szabó and Hédl 2011; Szabó 2015). Indeed, it can be argued that implicit
in the very study of ecology is a historical dimension (Figure 3.1), as ecologists
often seek to predict future responses of ecological systems, further predicting
spatial variation in their structure and function. This has become increasingly
relevant as ecologists consider ecosystem response to global change (Gilliam 2016).
Three decades of research at the Fernow Experimental Forest as part of the
Watershed Acidification Study underscores the essentiality of maintaining field
studies over long-term periods of time to address ecological processes, especially in
the context of global change.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Long-term Ecological Field Studies as Historical Ecology 41

3.5. References

Aber, J., McDowell, W., Nadelhoffer, K., Magill, A., Berntson, G., Kamakea, M., McNulty,
S., Currie, W., Rustad, L., Fernandez, I. (1998). Nitrogen saturation in temperate forest
ecosystems – Hypotheses revisited. BioScience, 48, 921–934.
Adams, M.B., De Walle, D.R., Hom, J. (2006). The Fernow Watershed Acidification Study.
Springer, Dordrecht, The Netherlands.
Chapman, S.K., Devine, K.A., Curran, C., Jones, R.O., Gilliam, F.S. (2016). Impacts of soil
nitrogen and carbon additions on forest understory communities with a high-deposition
history. Ecosystems, 19, 142–154.
Cholewińska, O., Adamowski, W., Jaroszewicz, B. (2020). Homogenization of temperate
mixed deciduous forests in Białowieża Forest: Similar communities are becoming more
similar. Forests, 11, 545.
Christ, M.J., Peterjohn, W.T., Cumming, J.R., Adams, M.B. (2002). Nitrification potentials
and landscape, soil and vegetation characteristics in two Central Appalachian watersheds
differing in NO3- export. For. Ecol. Manage., 159, 145–158.
Du, E. (2016). Rise and fall of nitrogen deposition in the United States. Proc. Natl. Acad. Sci.
USA, 113, E3594–E3595.
Du, E., Fenn, M.E., De Vries, W. (2019). Atmospheric nitrogen deposition to global forests:
Status, impacts and management option. Environ. Pollut., 250, 1044–1048.
Gilliam, F.S. (2006). Response of the herbaceous layer of forest ecosystems to excess
nitrogen deposition. J. Ecol., 94, 1176–1191.
Gilliam, F.S. (2016). Forest ecosystems of temperate climatic regions: From ancient use to
climate change. New Phytol., 212, 871–887.
Gilliam, F.S., Turrill, N.L., Aulick, S.D., Evans, D.K., Adams, M.B. (1994). Herbaceous layer
and soil response to experimental acidification in a central Appalachian hardwood forest.
J. Environ. Qual., 23, 835–844.
Gilliam, F.S., Yurish, B.M., Adams, M.B. (2001). Temporal and spatial variation of nitrogen
transformations in nitrogen-saturated soils of a central Appalachian hardwood forest. Can.
J. For. Res., 31, 1768–1785.
Gilliam, F.S., Hockenberry, A.W., Adams, M.B. (2006). Effects of atmospheric nitrogen
deposition on the herbaceous layer of a central Appalachian hardwood forest. J. Torrey
Bot. Soc., 133, 240–254.
Gilliam, F.S., Welch, N.T., Phillips, A.H., Billmyer, J.H., Peterjohn, W.T., Fowler, Z.K.,
Walter, C.A., Burnham, M.B., May, J.D., Adams, M.B. (2016). Twenty-five year
response of the herbaceous layer of a temperate hardwood forest to elevated nitrogen
deposition. Ecosphere, 7, e01250. 10.1002/ecs2.1250.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
42 Historical Ecology

Gilliam, F.S., Walter, C.A., Adams, M.B., Peterjohn, W.T. (2018). Nitrogen (N) dynamics in
the mineral soil of a central Appalachian hardwood forest during a quarter century of
whole-watershed N additions. Ecosystems, 21, 1489–1504.
Gilliam, F.S., Burns, D.A., Driscoll, C.T., Frey, S.D., Lovett, G.M., Watmough, S.A. (2019).
Decreased atmospheric nitrogen deposition in eastern North America: Predicted responses
of forest ecosystems. Environ. Pollut., 244, 560–574.
Gilliam, F.S., Adams, M.B., Peterjohn, W.T. (2020). Response of soil fertility to 25 years of
experimental acidification in a temperate hardwood forest. J. Environ. Qual., 49,
961–972.
Gress, S.E., Nichols, T.D., Northcraft, C.C., Peterjohn, W.T. (2007). Nutrient limitation in
soils exhibiting differing nitrogen availabilities: What lies beyond nitrogen saturation?
Ecology, 88, 119–130.
Groffman, P.M., Driscoll, C.T., Durán, J., Campbell, J.L., Christenson, L.M., Fahey, T.J.,
Fisk, M.C., Fuss, C., Likens, G.E., Lovett, G. et al. (2018). Nitrogen oligotrophication in
northern hardwood forests. Biogeochemistry, 141, 523–539.
Hurd, T.M., Brach A.R., Raynal D.J. (1998). Response of understory vegetation of
Adirondack forests to nitrogen additions. Can. J. For. Res., 28, 799–807.
Hutchings, M.J., John, E.A., Wijesinghe, D.K. (2003). Toward understanding the
consequences of soil heterogeneity for plant populations and communities. Ecology, 84,
2322–2334.
Levin, S.A. (1992). The problem of pattern and scale in ecology. Ecology, 73, 1943–1967.
Likens, G.E. (1989). Long-Term Studies in Ecology: Approaches and Alternatives.
Springer-Verlag, New York.
Lloret, J. and Valiela, I. (2016). Unprecedented decrease in deposition of nitrogen oxides over
North America: The relative effects of emission controls and prevailing airmass
trajectories. Biogeochemistry, 129, 165–180.
Lu, X., Mo, J., Gilliam, F.S., Zhou, G., Fang, Y. (2010). Effects of experimental nitrogen
deposition on plant diversity in an old-growth tropical forest. Glob. Change Biol., 16,
2688–2700.
May, J.D., Burdette, E., Gilliam, F.S., Adams, M.B. (2005). Interspecific divergence in foliar
nutrient dynamics and stem growth in a temperate forest in response to chronic nitrogen
inputs. Can. J. For. Res., 35, 1023–1030.
Peterjohn, W.T., Adams, M.B., Gilliam, F.S. (1996). Symptoms of nitrogen saturation in two
central Appalachian hardwood forests. Biogeochemistry, 35, 507–522.
Peterjohn, W.T., McGervey, R.J., Sexstone, A.J., Christ, M.J., Foster, C.J., Adams, M.B.
(1998). Nitrous oxide production in two forested watersheds exhibiting symptoms of
nitrogen saturation. Can. J. For. Res., 28, 1723–1732.
Richardson, H.L. (1938). The nitrogen cycle in grassland soils: With special reference to the
Rothamsted Park Grass Experiment. J. Agr. Sci., 28, 73–121.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Long-term Ecological Field Studies as Historical Ecology 43

Schmitz, A., Sanders, T., Bolte, A., Bussotti, F., Dirnböck, T., Johnson, J., Peñuelas, J.,
Pollastrini, M., Prescher, A.-K., Sardans, J. et al. (2019). Responses of forest ecosystems
in Europe to decreasing nitrogen deposition. Environ. Pollut., 244, 980–994.
Silvertown, J., Poulton, P., Johnson, A.E., Edwards, G., Heard, M., Biss, P.M. (2006). The
park grass experiment 1865–2006: Its contribution to ecology. J. Ecol., 94, 801–814.
Smith, H.C. and Miller, G.W. (1987). Managing Appalachian hardwood stands using four
regeneration practices, 34- year results. North. J. Appl. For., 4, 180–185.
Stoddard, J.L. (1994). Long-term changes in watershed retention of nitrogen: Its causes and
aquatic consequences. In Environmental Chemistry of Lakes and Reservoirs, Baker, L.A.
(ed.). American Chemical Society, Washington, DC, USA.
Szabó, P. (2015). Historical ecology: Past, present and future. Biol. Rev., 90, 997–1014.
Szabó, P. and Hédl, R. (2011). Advancing the integration of history and ecology for
conservation. Conserv. Biol., 25, 680–687.
Walter, C.A., Raiff, D.T., Burnham, M.B., Gilliam, F.S., Adams, M.B., Peterjohn, W.T.
(2016). Nitrogen fertilization interacts with light to increase Rubus spp. cover in a
temperate forest. Plant Ecol., 217, 421–430.
Walter, C.A., Adams, M.B., Gilliam, F.S., Peterjohn, W.T. (2017). Non-random species loss
in a forest herbaceous layer following nitrogen addition. Ecology, 98, 2322–2332.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
4

Gaps and Cracks in Land Cover


Mapping for Historical Ecology
Francesca Di PIETRO1, Roger COLY1, Clémence CHAUDRON2,
Samuel LETURCQ1
1
University of Tours, France
2
University of Lorraine, Nancy, France

4.1. Introduction

Land cover mapping at different historical periods has been a crucial step in
historical ecology studies, as it allows the identification of past land cover and the
quantification of land cover change, which have an influence on current ecological
habitats. Moreover, past land cover, used as a proxy of past land-use, is shown to
have an effect on the current composition of plant communities. Indeed, several
authors have shown that past rather than present landscape structure can strongly
affect present plant community composition (e.g. Lindborg and Eriksson 2004).
With regard to land cover history, geographical research on landscape changes,
which analyzes changes in land cover over time, has provided a significant
contribution. Extensive and accurate land cover mapping, allowing the collection of
spatial information on the whole gradient of land cover classes, is particularly
valuable. Nevertheless, a large part of historical ecological studies is based on the
mapping of one land cover class (often forest patches), mainly based on topographic
maps (e.g. Kaim et al. 2016). Despite this limitation, studies on non-forest elements,
which require a more detailed knowledge of agricultural land cover, with specific
information on both arable land and permanent grassland patches, are growing (e.g.
Skaloš and Engstovà 2010). Yet, two types of difficulties for extensive land cover
mapping hinder the development of studies in historical ecology: (i) land cover

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
46 Historical Ecology

mapping at different historical periods is highly time-consuming and relatively


uncertain, as there are few possibilities for checking the accuracy of data sources;
(ii) comparison of spatial information from different historical sources, including
modern, accurate and detailed databases, leads to a loss of precision, as the less
precise source determines the final land cover categorization. These difficulties
depend on the type of data available over the period studied. Several authors, mainly
geographers and historians, indicate some turning points of land-use change in the
modern era, documented in several European countries. Some successive turning
points can be defined in rural and urban landscapes evolution (Antrop 2005). (1) In
the 19th century, the first part of the century represents the peak in population and
therefore in land-use pressure in rural areas (Bergès and Dupouey 2017), as well as
the rise of urbanization in the industrial era in urban areas. (2) In the 20th century,
the 1950s/1960s represent the beginning of extensive mechanization and
intensification of agriculture in rural areas (Meeus et al. 1990), as well as post-war
reconstruction in urban areas. The 1970s/1980s represent the moment when the
effects of the common agricultural policy (CAP) can be observed in the rural
territory, particularly in terms of landscape homogeneity, and the period of suburban
extension in cities. In this chapter, we address both main constraints of historical
ecology studies: the use of historical maps and the comparison of spatial information
from different historical sources.. We address particularly land cover mapping of the
19th century, the 1970s/1980s, and the present period.

4.2. Three main steps of past land cover mapping

Mapping past land cover with geographical information system software


generally encompasses three steps: 1) the georeferencing of past maps, which allows
the accurate location of images by assigning their geographical coordinates on a
georeferenced vector layer; (2) the vectorization of land cover patches and linear
elements (e.g. roads and watercourses); and (3) the classification of land cover
patches, i.e. the attribution of a specific land cover class to each polygon. The
automation of the last two steps is crucial to reduce the time of mapping and could
help the study of large areas in a precise way, i.e. by considering the whole range of
land cover classes and in particular by distinguishing, in agricultural areas,
permanent grasslands from arable lands. The graphic quality of the images
influences the probability of successfully automating land cover mapping. Two
elements deserve particular attention: (i) the contrast of the colors indicating the
heterogeneity of land cover patches and allowing both the vectorization and
classification of the land cover patches and (ii) maps should be free of written
indications, as these prevent vectorisation. If these two conditions are met, automatic
vectorization and classification are possible, by using different methods of image
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Gaps and Cracks in Land Cover Mapping for Historical Ecology 47

processing and software (e.g. CorelDRAW, ArcScan, an extension of ArcMap


or HistMapR package). Otherwise, if the maps’ graphic quality does not allow
automatic vectorization, manual vectorization is necessary. A slightly less
time-consuming and more accurate alternative is to draw the parcel grid manually on
tracing paper and then proceed with digitalization, georeferencing and automatic
vectorization. The classification, which follows this step in all cases, is then based
on visual interpretation of aerial photographs (e.g. Skaloš and Engstovà 2010) or on
written information in old cadastral registers (e.g. Skaloš and Kašparová 2012).

4.3. Land cover in the 19th century: the old cadasters

In most European countries, cadasters giving an accurate information on land


cover at the parcel scale were produced in the 19th century for fiscal purposes (see
Box 4.1). The term “parcel” refers to “any portion of land not divided by material
separations, having the same type of culture and belonging to the same owner”
(Herbin and Pebereau 1953 in Bartout 2011).

From the end of the 17th century, a great movement of fiscal reforms began in
Europe, in order to resolve the financial deficits linked to the wars in which almost all the
states of the continent were engaged. These reforms were directed towards a system of
direct taxation: the cadasters. The exceptions were for Switzerland, which had renounced
waging external wars at the beginning of the 16th century, and England, which turned
towards indirect taxes, due to the opposition of the nobility to direct taxation. During the
18th and 19th centuries, the first cadasters were carried out, based on owners’ declarations
(as in Luxembourg and Spain), or based on geometrical parcel measurements (as in the
Duchy of Milan). The latter, i.e. the parcel cadasters, allowed all parcels to be classified
according to the degree of soil fertility. In France, after an unsuccessful attempt in 1763,
the parcel cadaster was finally created after the Revolution, by a decision of Emperor
Napoleon in 1807. Classifying 100 million parcels, the “Napoleonic cadaster” was
completed in more than 40 years (Clergeot 2007).

Box 4.1. The genesis of national cadastral policies in Europe

Old cadasters are used in historical ecological studies, particularly in central


Europe, northern Europe and in Mediterranean areas. Parcel maps drawn during the
previous periods, mainly the 17th and 18th centuries, are sometimes available and
used in ecological historical studies, as in Sweden (e.g. Cousins 2001). Before these
periods, no systematic parcel maps are available at a national scale (e.g. Szabo and
Hédl 2011). This is different for local scales (see Box 4.2).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
48 Historical Ecology

It was in the 13th century that fiscal documentation was produced to document the
land-use in a given territory by describing the parcels that constituted the properties.
These were either manorial sources of a private character used to manage the feudal lord’s
estate (Brunel et al. 2002), or sources of a public character (called “cadastres”, “estimes”,
“compoix”, etc.) which served, in Mediterranean Europe, to raise taxes (Abbé et al. 2017).
It was not until the end of the Middle Ages that the first parcel plans were produced in
support of these tax lists. These plans, which complemented the registers, became
increasingly common in the 17th and 18th centuries in Europe.

Box 4.2. The genesis of land-plot mapping in Europe

The main benefit of old cadasters is that they include the whole range of land
cover classes, and not only a few of them, such as forested, urban and water patches,
which are included in topographic maps and often used in historical ecological
studies focusing on forest landscapes. Moreover, because of the fiscal reasons for
their creation, they are known to be accurate and precise (e.g. Rochel et al. 2017).
Nevertheless, their use is highly time-consuming. Two main types of material
constitute most of the old cadasters (Figure 4.1).

Figure 4.1. Napoleonic cadaster (1808–1832), municipality of Nouans-les-Fontaines.


Joining map (a), parcel map (b), detail of a parcel (c) and parcel register (d) (source:
Archives of the Indre-et-Loire department, France). For a color version of this figure,
see www.iste.co.uk/decocq/ecology.zip
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
49

department, France). For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip


Gaps and Cracks in Land Cover Mapping for Historical Ecology

(b), to get land cover specification (c) and mapping (d) (source: Archives of the Indre-et-Loire
Figure 4.2. Processing of Napoleonic cadaster (1808–1832). Parcel map (a), compared to parcel register
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
50 Historical Ecology

(i) Parcel maps: in France, parcel maps are drawn up on a scale that varies
according to the density of the elements represented (from 1/500 for urban areas to
1/5,000 for rural areas); these digitized parcel maps (raster format) are available on
the website of the department archives.

(ii) Parcel registers, which mention the type of land cover of each parcel.

After georeferencing and manual vectorization of the parcel maps, classification


of land cover patches is required. This step consists of assigning to each parcel
polygon the land cover type from the information given on the cadastral registers.
Indeed, the graphic quality of parcel maps does not allow for a direct reading of land
cover and automatic mapping of patches. This is due to the low color contrast
between land cover types, on the one hand, and to the presence of written
information on the maps, with a variety of details on land cover or other elements,
on the other hand (Figure 4.2). In order to optimize the cartographic representation,
adjacent parcels with the same land cover class can be merged (Figure 4.3).

Figure 4.3. Vectorization of Napoleonic cadaster (1808–1832) by merging parcels


with the same land cover type (a, b). “Terre” (Arable) Land, “Bruyère”: Heather.
For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip

4.4. Land cover in the 20th century: aerial photographs

The use of vertical aerial photographs to identify land cover in various periods of
the 20th century is common in historical ecology and landscape studies (see Gerard
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Gaps and Cracks in Land Cover Mapping for Historical Ecology 51

et al. 2010). Exhaustive photographic land surveys on a large spatial scale and at
regular time intervals were carried out in a majority of European countries and
elsewhere, first with black and white photographs (from the 1940s onwards) and
then with color photographs (from the 1990s onwards). Moreover, near-infrared
photographs, often used in forest-related studies, allow precise identification of
chlorophyll activity (color infrared) and wetlands (black and white infrared). The
digitized images are often freely available on governmental websites, allowing a
wide spatial scope. One of the main interests of aerial photographs is the possibility
they offer to identify land cover at several dates, with regular and close dates, and
thus capture its changes throughout the 20th century. In our study (Chaudron et al.,
2018), the methodology of land cover mapping is based on black and white aerial
photographs for the years 1978 and 1980 from National Geographic Institute
databases. Each land cover patch is outlined using a semi-automatic vectorization
method with CorelDRAW software, and visually checked. Thus, after
georeferencing, the semi-automatic vectorization is carried out through two steps
(Figure 4.4): (1) an initial vectorization using an image processing software
(CorelDRAW v7); (2) then, after a new georeferencing of the vectorized image
using the already georeferenced photo, corrections of the vectorized image were
necessary to (i) avoid overlaps and gaps, (ii) merge unfairly divided plots or split
unfairly aggregated plots, (iii) eliminate polygons that are too small by grouping
polygons with the same type of land cover: a threshold of 0.5 ha was chosen to
simplify the operation without losing much surface area and (iv) integrate linear
elements (road and river networks etc.). The past land cover was then defined based
on visual photointerpretation.

We observe that the process of semi-automatic cartography includes, in each of


the two main steps, vectorization and photointerpretation, two kinds of difficulties,
which limit their use. First, the vectorization produces some extra elements (lines) or
deficient elements (gaps) as well as overlaps within and between polygons;
moreover, as the distinction between different crops is not always clear on the photo,
manual corrections are necessary in order to increase the precision of parcel edges.
Second, the photointerpretation of land cover is hampered by the uncertainty of the
attribution of a polygon to a land cover class, as the contrast of colors is not always
intense and clearly legible. In particular, if the aerial photographs used are not
digitized, but are printed on paper, nuances of photographic development and
subsequent colors can change from one photo to another: this heterogeneity creates
artificial differences and prevents correct vectorization and photointerpretation.
Although it is easier to automate, as color contrast between land cover patches is
higher than in old cadasters’ parcel maps, and there are no written indications on the
image, the use of this source of information on land cover seems paradoxically less
reliable than the old cadasters, which clearly indicate land cover on parcel registers.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
52 Historical Ecology

Figure 4.4. Processing of aerial photographs (date: 1978). Top: vectorization;


georeferenced photo (a), first vectorization using an image processing software (b),
new georeferencing of the vectorized image using the already georeferenced photo
(c). Bottom: corrections of the vectorized image. Unfairly divided plots (d), merging
unfairly divided plots (e), unfairly aggregated plots (f) and splitting unfairly aggregated
plots (g). For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip

4.5. Present land cover: modern databases

In combination with past land cover mapping, current land cover mapping is
often used in historical ecology studies to measure changes in land cover over time
and to compare current and past land cover in terms of their relationship to
biodiversity. Contrary to other historical periods, the current period benefits, for land
cover mapping, from databases that are already georeferenced, vectorized and
informed, which avoid using the three main steps of past land cover mapping
mentioned above (georeferencing, vectorization, classification). However, to obtain
complete and detailed information, not limited to the main land cover classes such as
forested, urban and aquatic areas, it is necessary to integrate into a geographic
information system a set of databases from different sources. In particular,
information on agricultural land must be added to information on the main land
cover classes (forested, urban and aquatic areas) generally available on government
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Gaps and Cracks in Land Cover Mapping for Historical Ecology 53

websites (in France, this includes the National Geographic Institute’s BDTopo®). In
France, localized information on agricultural land cover is gathered in the graphic
parcels register (GPR, “Registre Parcellaire Graphique”), a database on crops on
localized parcels, which is derived from annual declarations made by farmers in
respect of the CAP. This source of information, which is made available annually by
the Ministry of Agriculture in open access, is valuable for its accuracy on crops, but
it requires two types of corrections. On the one hand, information on crop type is
available for the main crop of a crop block (set of adjacent parcels used by one
farmer) and not at the parcel scale itself: only the main crop per block is shown. It is
therefore often necessary to split the crop block into several parcels, by using
cadastral digitized maps and by visually checking the orthorectified digital
photographs, to obtain information on land cover at the parcel scale. On the other
hand, there is missing information about the crop types of some parcels, as some
crops are unreported to the CAP (generally permanent grassland or vines). The
photointerpretation of the orthophotos is therefore essential to fill these gaps. For
both the corrections, it is therefore essential that the GPR and the orthophotos be of
the same year. In addition, it should be noted that this crop information, intended for
the payment of agricultural subsidies, is generally excessively detailed for historical
ecology studies (e.g. the GPR distinguishes winter barley crops from winter wheat
crops, etc.). The aggregation of crop classes may be necessary, particularly when
unifying the final land cover classification among several periods and sources (see
below). Finally, the compatibility between these two main databases (general land
cover and specifically agricultural land cover) must be ensured by choosing a
compatible date, and by excluding possible overlap between polygons from different
databases. Alternatively, maps derived from remote sensing of satellite images are
available, such as the maps of the Biosphere’s Centre for Spatial Studies, based on
Landsat-8 data (Inglada et al. 2017). These raster images can be easily vectorized
and their photointerpretation refined by the databases mentioned above.

4.6. From different sources to one land cover typology

From the beginning of the investigations, it is recommended to establish a single


land cover classification compatible with the different sources. Indeed, after having
mapped the land cover at different dates and from different sources, it is necessary
to unify the information on land cover by a single nomenclature. This step, which
necessarily leads to a loss of information compared to the most precise information
(generally the most recent), is nevertheless essential. It involves compromises about
precision, as the less precise source of land cover information leads to the final
categorization. It also involves simplification of land cover information, and
decrease in the number of unified land cover classes. It should be noted that the
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
54 Historical Ecology

parcel registers of the old cadasters include fairly detailed land cover classes
(Table 4.1), allowing all parcels to be classified according to the degree of soil
fertility for that period, making the establishment of fiscal rules possible. It is worth
noting, for example, the plurality of precise terms used to indicate permanent
grassland (“Meadow”, “Pasture”1), semi-natural areas (“Heathland”, “Moorland”,
“Marshes”) and uncultivated areas (“Wasteland”, “Fallow land” or “terre vaine”).
The latter term, “terre vaine”, indicates an arable land which is not sown at the time
of the cadastral survey and may be widely open to grazing by village herds if the
specific range pasturing system (“vaine pâture”) is applied. Each land cover class
represented at that time a specific level of soil fertility and subsequent income,
which is not immediately clear today, especially since some classes have almost
disappeared (e.g. the classes “Planted pasture”2 and “Hemp”). These classifications
must be understood according to the uses implemented during the specific historical
period studied, in order to integrate them correctly into a diachronic classification,
which requires historical information on past agricultural practices.

Land cover
CODE Land cover class CODE Land cover class CODE
class
1 Arable land 8 Hemp 15 Wasteland
2 Meadows 9 Heathland 16 Pond
3 Garden (kitchen 10 Moorland 17 Ditch
garden)
4 Vine 11 Marshes (flooded area) 18 Building
5 Orchard 12 Tree nursery 19 Cemetery
6 Planted pasture 13 Forest 20 Quarry
7 Pasture 14 Fallow land (terre
vaine)

Table 4.1. Land cover classes for the parcels of the


study site, Napoleonic cadaster (1808–1832)

1. The terms “meadow” and “pasture” refer to grazed grassland, but the distinction between
the two terms shows a nuance that is difficult to define for the past periods. It is possible that
the word “meadow” refers to a sown and possibly mowed pasture, while the word “pasture”
refers to a grazed area where spontaneous vegetation grows. See Brumont, F. (2008). Prés et
pâtures en Europe occidentale. Presses universitaires du Mirail, Toulouse.
2. Planted pasture probably refers to a grassy area planted with fruit trees, used as pasture for
livestock.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Gaps and Cracks in Land Cover Mapping for Historical Ecology 55

4.7. Conclusion

The acquisition of localized and accurate information on land cover over time is
a crucial step in understanding not only landscape changes, but also changes in
ecological patterns, and the legacy that past landscapes represent for present
biodiversity. This step is time-consuming, and semi-automatic mapping methods are
still in their early stages; they are uncertain and require careful adaptation to the
specificities of the information sources. This explains why studies in historical
ecology are somewhat limited, being confined to small areas, and restricted to coarse
land cover categories (e.g. urban, forested, aquatic, agricultural areas). This
restriction does not allow the identification of key categories such as permanent
grasslands, indiscriminately associated with arable land in a broad category of
“agricultural land”. However, semi-automatic vectorization and photointerpretation
methods are developing considerably and are leading to the emergence of the field
of historical geographical information systems. In the near future, this would
undoubtedly help overcome the current hurdles in historical land cover mapping and
help to achieve large-scale studies in historical ecology.

4.8. References

Abbé, J.-L., Hautefeuille, F., Jaudon, B., Le Pottier, J., Olivier, S. (2017). Estimes, compoix
et cadastres. Histoire d’un patrimoine commun de l’Europe méridionale. Le Pas d’Oiseau,
Toulouse.
Antrop, M. (2005). Why landscapes of the past are important for the future. Landscape and
Urban Planning, 70(1–2), 21–34 [Online]. Available at: https://doi.org/10.1016/j.
Bartout, P. (2011). L’apport du cadastre napoléonien aux problématiques spatiales des
retenues d’eau. Revue géographique de l’Est, 51(3–4). doi: https://doi.org/10.4000/
rge.3382.
Bergès, L. and Dupouey, J. (2017). Écologie historique et ancienneté de l’état boisé :
concepts, avancées et perspectives de la recherche. Revue forestière française, 4–5(1),
297–318.
Brumont F. (2008) Prés et pâtures en Europe occidentale. Presses universitaires du Mirail,
Toulouse.
Brunel, G., Guyotjeannin, O., Moriceau, J.-M. (2002). Terriers et plans-terriers du XIIIe au
e
XVIII siècle. Actes du Colloque de Paris (23–25 septembre 1998). Association d’Histoire
des sociétés rurales/École nationale des Chartes, Caen-Paris.
Chaudron, C., Perronne, R., Bonthoux, S., Di Pietro, F. (2018). A stronger influence of past
than present landscape structure on plant communities of road-field boundary. Acta
Oecologica, 92(November 2017), 85–94. doi: 10.1016/j.actao.2018.08.009.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
56 Historical Ecology

Clergeot, P. (2007). 1807 – Un cadastre pour l’Empire. Cent millions de parcelles en France.
Publi-Tope, Paris.
Cousins, S.A.O. (2001). Analysis of land-cover transitions based on 17th and 18th century
cadastral maps and aerial photographs. Landscape Ecology, 16(1), 41–54. doi: 10.1023/A:
1008108704358.
Gerard, F., Petit, S., Smith, G., Thomson, A., Brown, N., Manchester, S., Wadsworth, R.,
Bugar, G., Halada, L., Bezak, P. et al. (2010). Land cover change in Europe between 1950
and 2000 determined employing aerial photography. Progress in Physical Geography,
34(2), 183–205.
Inglada, J., Vincent, A., Arias, M., Tardy, B., Morin, D., Rodes, I. (2017). Operational high
resolution land cover map image time series. Remote Sensing, 9(95), 1–35. doi:
10.3390/rs9010095.
Kaim, D., Kozak, J., Kolecka, N., Ziółkowska, E., Ostafin, K., Ostapowicz, K., Gimmi, U.,
Munteanu, C., Radeloff, V.C. (2016). Broad scale forest cover reconstruction from
historical topographic maps. Applied Geography, 67, 39–48. doi: 10.1016/
j.apgeog.2015.12.003.
Lindborg, R. and Eriksson, O. (2004). Historical landscape connectivity affects present plant
species diversity. Ecology, 85(7), 1840–1845.
Meeus, J.H.A., Wijermans, M.P., Vroom, M.J. (1990). Agricultural landscapes in Europe and
their transformation. Landscape and Urban Planning, 18(3), 289–352.
Rochel, X., Avon, C., Bergès, L., Chauchard, S., Grel, A. (2017). Quelles sources
cartographiques pour la définition des usages anciens du sol en France ? Revue forestière
française, LXIX(4–5), 353–370 [Online]. Available at: https://doi.org/10.4267/2042/
67866.
Skaloš, J. and Engstovà, B. (2010). Methodology for mapping non-forest wood elements
using historic cadastral maps and aerial photographs as a basis for management. Journal
of Environmental Management, 91(4), 831–843. doi: 10.1016/j.jenvman.2009.10.013.
Skaloš, J. and Kasparová, I. (2012). Landscape memory and landscape change in relation to
mining. Ecological Engineering, 43, 60–69. doi: 10.1016/j.ecoleng.2011.07.001.
Szabo, P. and Hedl, R. (2011). Advancing the integration of history and ecology.
Conservation Biology, 25(4), 680–687. doi: 10.1111/j.1523-1739.2011.01710.x.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
5

The Use of Repeat


Photography in African
Historical Ecology
Michael Timm HOFFMAN and Rick F. ROHDE
University of Cape Town, South Africa

5.1. Repeat photography as an emerging tool in African historical


ecology

One goal of historical ecology is to reconstruct past ecological patterns and


dynamics (Beller et al. 2020). Knowing how environments have changed over time
in response to key drivers and processes helps us to better understand and manage
the present. Describing the trajectory in terms of the nature, extent and rate of past
change can also indicate what to expect in the future. Repeat photography is a useful
tool in the exploration of this past–present–future environmental change continuum
(Hoffman et al. 2019).

Repeat photography involves standing on the same location from where an


earlier photograph was made and retaking the image as closely matched to the
original photograph as possible (Webb et al. 2010). On-site field surveys and
detailed post-production analyses of the matched repeat photographs can provide
an accurate account of long-term change at a location (Hammond et al. 2020).
Observed variations in species and growth forms (e.g. trees, shrubs and grasses) in
different parts of the landscape (e.g. slopes, plains, riverine areas) provide additional
information about the pattern of change and the kind of pressures that have been
placed on the environment over time. If the trajectories observed in the repeat

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
58 Historical Ecology

photograph pairs are replicated at many different locations, then regional (Jacob
et al. 2017) or national (Hoffman et al. 2019) syntheses of long-term environmental
change are sometimes possible.

Repeat photography has been used globally for decades to evaluate landscape
change (Rogers et al. 1984). Even though one of the earliest applications of repeat
photography was undertaken in Africa (Shantz and Turner 1958), it has only been in
the last few decades that the technique has been more widely applied on the
continent. Nearly three quarters of the 60 main studies which use repeat
photography to understand environmental change in Africa have been published
since 2010. This chapter documents where these studies have been concentrated on
the continent and the main themes that have been addressed using this approach. It
also describes the strengths and limitations of the technique and points to some
promising future directions.

5.2. Repeat photography and landscape change in Africa

5.2.1. Early contributions

Shantz and Turner’s (1958) account of vegetation change in Africa is one the
earliest uses of repeat photography. It documents long-term changes in vegetation at
two time steps (1919 and 1956) along a 4,000 km transect from Cape Town to
Nairobi. An analysis of repeat photographs was only used again three decades later
by Hoffman and Cowling (1990). Matched photographs of the eastern Karoo region
of South Africa were used to assess the hypothesis of desertification that had been
proposed earlier by Acocks (1953). Since then, the number of studies which have
used repeat photography to understand long-term environmental change on the
continent have increased in both number and scope. Isolated studies have been
undertaken in several different countries including in Madagascar (Kull 2005),
Senegal (Herrmann and Tappan 2013), Kenya (Turner et al. 1998; Western 2010),
Tanzania (Rohde and Hilhorst 2001) and Eritrea (Lätt 2004). However, there has
been a concentration of repeat photography studies in Ethiopia and southern Africa.

5.2.2. Ethiopia

Most of the research in Ethiopia has been undertaken in the semi-arid, northern
highlands of Tigray by a group of geographers at the University of Ghent in
Belgium (e.g. Nyssen et al. 2008, 2014; De Mûelenaere et al. 2014; Jacob et al.
2015). A major focus has been on assessing the extent of forest cover in early
colonial environments and how this has been affected by anthropogenic influences
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Use of Repeat Photography in African Historical Ecology 59

(Meire et al. 2013). The role of protected areas in preventing deforestation in the
Simien Mountains has also been evaluated using historical photographs as has the
upward shift in the treeline in areas protected from high anthropogenic pressure
(Jacob et al. 2017). Significant effort has been expended on land management
interventions in the region and repeat photography has been used to document the
impact of these interventions on vegetation cover and soil erosion (Frankl et al.
2012, 2015; Nyssen et al. 2015). Historical photographs have also been used in
combination with an analysis of fossil pollen from the endorheic Lake Ashenge in
North Ethiopia to understand the cyclical nature of deforestation and forest regrowth
over the past 400 years (Lanckriet et al. 2015).

5.2.3. Southern Africa

Repeat photography studies in southern Africa have focused on Namibia and


South Africa. A major emphasis has been on evaluating the regional climate model
forecasts which project widespread aridification for the subcontinent in the future.
Historical trajectories of ecological patterns and dynamics have been used to assess
the direction and magnitude of change suggested in the model outputs. An analysis
of 52 photographs originally taken in 1876 along a 1,000 km transect in Namibia
showed how both rainfall and land-use affect vegetation (Rohde and Hoffman 2012).
Sites with annual rainfall below 250 mm were relatively stable over time, while
woody cover had increased significantly at locations with higher annual rainfall
totals. An analysis of 100 repeat photographs from the Namib desert provides
evidence of climate change related to regional synoptic changes in savanna rainfall
and the fogbelt associated with the coastal Benguela upwelling system (Rohde et al.
2019).

In South Africa, long-term changes in vegetation cover in most of the major


biomes have been investigated using repeat photography. In some localities in the
Fynbos biome, forests have expanded into areas that were previously dominated by
fynbos shrublands largely in response to the suppression and lower frequency of
fires relative to the 19th century fire regime (Poulsen and Hoffman 2015). The
semi-arid, species-rich Succulent Karoo biome has been relatively stable in terms of
vegetation cover on the plains and rocky slopes (Hoffman and Rohde 2007;
Figure 5.1), while riverine areas have experienced an increase in the cover of woody
plants (Hoffman and Rohde 2011). In the eastern Nama–Karoo biome, there has
been an increase in grass cover, especially over the last 50 years, in response to an
increase in rainfall and a decrease in livestock (Masubelele et al. 2015; Figure 5.2).
The Savanna biome has experienced the greatest change in vegetation cover over the
20th century. In both mesic and semi-arid locations, the cover of woody plants has
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
60 Historical Ecology

increased significantly (Hoffman and O’Connor 1999; Russell and Ward 2014;
Ward et al. 2014; Figure 5.3). In some cases, the increase in woody plant cover has
occurred over just a few decades in natural environments as well as on abandoned
croplands. Attribution for these changes has been difficult, and a range of drivers
have been invoked to explain the increase in woody plants including changes in
herbivory, fire, land-use, rainfall, temperature and CO2 (O’Connor et al. 2014).

Figure 5.1. Most matched photographs of the Succulent Karoo biome show stability
in vegetation cover over time. Top: Pole Evans c. 1920. Bottom: Rohde and Hoffman,
October 27, 2006. Calvinia (No. 246). For a color version of this figure, see
www.iste.co.uk/decocq/ecology.zip

A recent analysis of over 1,300 repeat historical photographs of the biomes of


southern Africa provides a summary of the main changes that have taken place in
the vegetation of the region over the last 100 years (Hoffman et al. 2019). The
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Use of Repeat Photography in African Historical Ecology 61

long-term trends suggest that the direction of vegetation change is different from
that which is expected, based on the outputs from most model forecasts. The
trajectories evident in most repeat photograph comparisons reflect environments that
have not yet been impacted by climate change but rather landscapes that are
recovering from the heavy exploitation of the 19th and early 20th centuries.

Figure 5.2. At the ecotone between the Nama–Karoo and Grassland biomes, there
has been an increase in grass cover. Top: Acocks, July 25, 1946. Bottom: Hoffman
and Arena November 2018. Blesbokvlakte 2 (No. 1844). For a color version of this
figure, see www.iste.co.uk/decocq/ecology.zip
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
62 Historical Ecology

Figure 5.3. The mesic savannas have experienced a significant increase in woody
plant cover. Top: Edwards. October 24, 1955. Bottom: Puttick, April 19, 2011.
Weenen-Mooi River Road, KwaZulu-Natal (No. 108). For a color version of this
figure, see www.iste.co.uk/decocq/ecology.zip

5.3. Long-term change in plant populations as revealed by repeat


photography

Most studies have used an analysis of historical photographs to document


landscape level changes in the environment. A few studies, however, have also used
closely matched repeat photographs to document changes in plant populations over
time. For example, Duncan et al. (2006) counted the number of individuals of the
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Use of Repeat Photography in African Historical Ecology 63

desert tree aloe, Aloidendron pillansii, evident in 14 matched photographs from


1950 and 2006, to show a 1.4% annual mortality of the species in the Richtersveld,
South Africa. Explanations for the decline include age-related natural causes,
climate change, grazing and theft. Age-related annual growth rates of individual tree
aloes can also be determined from well-matched repeat photographs and used in
population models (Hoffman et al. 2010).

Figure 5.4. Repeat photographs can be used to assess long-term changes in some
species such as the population of Aloidendron dichotomum. There were 66
individuals in the original photograph and only eight in the repeat. Other species in
the image include the grass, Stipagrostis ciliaris, and the succulent shrub, Euphorbia
gummifera. Top: Marloth c. 1930s. Bottom: Hoffman and Jack, May 25, 2008. Garub,
Namibia (No. 472). For a color version of this figure, see www.iste.co.uk/decocq/
ecology.zip
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
64 Historical Ecology

Changes in the status of 1,313 live trees of the iconic conifer, Widdringtonia
cedarbergensis, in the Cederberg mountains in South Africa were also assessed
using 87 repeat photo pairs taken between 1931 and 2013 (White et al. 2016).
Counts suggest that 74% of the individuals had died and only 3.4% new individuals
had recruited into the population over the period of observation. A generalized linear
mixed-effects model (GLMM) suggested that several factors were related to the
mortality of this species including greater fire frequency and increased temperatures.
Individuals growing at lower elevations and in less rocky habitats were also more
susceptible to dying. A third example of how repeat photographs can be used to
document long-term changes in plant populations is that of Okubamichael et al.
(2016) who studied the fate of 626 individual cycads documented in 107 repeat
photographs at 53 locations in South Africa. Their results showed that by 2014, only
16% of the original cycads, evident in the photographs, had survived. They
attributed the decline primarily to the theft of individuals for the lucrative cycad
trade although the harvesting of bark for medicinal purposes was also important in
some areas.

The technique is clearly not suitable for all species. However, for those species
with individuals that can be easily counted in photographs (Figure 5.4), the use of
repeats can provide useful, quantitative evidence of long-term change in
populations. When coupled with a GIS analysis of the location of individuals in the
landscape, together with information of key drivers of change such as fire and
rainfall, relatively robust predictions about the future trajectory of plant populations
can be made (White et al. 2016).

5.4. Strengths and limitations

Repeat photography provides an accurate, practical tool to document the nature,


extent and rate of change in key elements of the landscape. When the patterns of
change are assessed together with other historical sources such as long-term fire,
herbivory and rainfall records, the relative influence of both natural and
anthropogenic impacts on the environment can be evaluated (Hoffman et al. 2020).
A detailed analysis of historical photographs also fills an important gap in the spatial
and temporal scales of observation for global change research (Hammond et al.
2020). Repeat landscape photographs cover a greater area than botanical survey plot
data but less area than that usually assessed in aerial photographs or satellite images.
While repeat photographs do not have the same level of details about species and
populations that are contained in plot data, the quality of this information is greater
than that which can be measured from aerial photographs and satellite images
(Venter et al. 2020). Perhaps, the most significant contribution that historical
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Use of Repeat Photography in African Historical Ecology 65

photographs make to an understanding of environmental change is that they often


extend over 100 years or more (Rohde and Hoffman 2012). This temporal depth is
greater than the other approaches mentioned above and helps to set baselines for
conservation against which management decisions can be evaluated (Gillson 2015;
Beller et al. 2020).

One limitation of repeat photography, and one which it shares with many other
historical studies, is the problem of “snapshot” resampling (Stuble et al. 2020). This
occurs when only two (or a few) time points are sampled over an extended period.
When only the start and end points are sampled, as is usually the case in repeat
photography studies, inter-annual variance in the response variable (e.g. plant cover,
erosion gulley depth) cannot be determined. Because of this, important signals of
long-term change are missed or spurious conclusions about the long-term trends in
the response variables are reached (McCain et al. 2016). This can be a significant
problem for determining the long-term trend in the populations of short-lived
species, but for long-lived species, the same individuals are often found in
photographs taken decades apart (Rohde and Hoffman 2012). Knowledge of the
longevity of species helps significantly in the interpretation of change observed in
repeat photographs as does an understanding of the dynamics and drivers of the
vegetation type under investigation (Hoffman et al. 2020).

On their own, a set of matched photographs has only limited value. Any change
observed between two photographs requires additional information in order to
understand both the context for the pattern and to provide an explanation for the
differences. The use of archival sources, traveler’s journals and climate and
population census data, together with detailed on-site field surveys, often helps with
this (Hoffman et al. 2020). A thorough ethnography of relevant land users and
institutions is also necessary to appreciate the wider political and economic aspects
of environmental change as well as the perspectives of local people (Davis 2009;
Von Hellerman 2020). However, attributing change to any one factor is difficult, as
it is often several, interacting causes which are thought to be responsible for the
differences observed between two photographs (Turner et al. 2003).

5.5. Future directions

The techniques associated with measuring changes in different components of


the landscape in matched photo pairs have remained relatively constant over the last
20 years. Most studies use a visual assessment of the relative changes in the pair of
photographs to determine long-term trends. Often, the assessment relies on an expert
rating system involving multiple observers (Jacob et al. 2017; Hoffman et al. 2019).
More objective approaches such as point frequency estimates of plant species cover
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
66 Historical Ecology

and number as well as automatic image analysis techniques have also been advanced
(Hoffman and Todd 2010). However, these latter techniques have not been adopted
widely perhaps because of the greater efficiency and accuracy of visual estimations
(Vanha-Majamaa et al. 2000). An objective, widely acceptable, rapid technique to
estimate the cover of different components in digital images is still needed.

Repeat photographs on their own are extremely useful for understanding


environmental change. However, ground-based images often only cover a limited
field of view. When used interactively with other methods such as aerial
photographs and remotely sensed images, repeat photographs can provide a
reference for metrics of vegetation change derived from these other sources, such as
Landsat data (Venter et al. 2020). The integration of historical photographs with
other long-term approaches to understand environmental change is only likely to
grow. Their value in testing some of the forecasts made by global change models
will also be more carefully considered in the future (Hoffman et al. 2020).

Finally, the rapid increase in the use of this technique across the continent
suggests that it has been widely adopted as an important tool in global change
research. Photographs provide a visually accessible account of long-term change
that can be understood and appreciated by a wide range of interested observers.
Potential for even greater use has been promoted via the development of citizen
science networks (Hammond et al. 2020). The success of rePhotoSA
(http://rephotosa.adu.org.za), a citizen science project which encourages members of
the public to download and retake historical landscape photographs of southern
Africa (Scott et al. 2018), suggests that a wider adoption of the technique is possible.
An expanded, easily accessible collection of repeat photographs provides
information of long-term environmental change and has obvious implications for
ecosystem management.

5.6. References

Acocks, J.P.H. (1953). Veld types of South Africa. Mem. Bot. Surv. Sth. Afr., 28, 1–192.
Beller, E.E., McClenachan, L., Zavaleta, E.S., Larsen, L.G. (2020). Past forward:
Recommendations from historical ecology for ecosystem management. Glob. Ecol.
Conserv., 21, e00836.
Davis, D.K. (2009). Historical political ecology: On the importance of looking back to move
forward. Geoforum, 40, 285–286.
De Mûelenaere, S., Frankl, A., Haile, M., Poesen, J., Deckers, J., Munro, N., Veraverbeke, S.,
Nyssen, J. (2014). Historical landscape photographs for calibration of landsat land
use/cover in the northern Ethiopian highlands. Land Degrad. Dev., 25, 319–335.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Use of Repeat Photography in African Historical Ecology 67

Duncan, J., Hoffman, M.T., Rohde, R.F., Powell, E., Hendricks, H. (2006). Long-term
population changes in the giant quiver tree, Aloe pillansii in the Richtersveld, South
Africa. Plant Ecol., 185, 73–84.
Frankl, A., Poesen, J., Deckers, J., Haile, M., Nyssen, J. (2012). Gully head retreat rates in the
semi-arid highlands of northern Ethiopia. Geomorphology, 173, 185–195.
Frankl, A., Poesen, J., Moeyersons, J., Nyssen, J. (2015). Gully development in the Tigray
highlands. In Landscapes and Landforms of Ethiopia, Billi, P. (ed.). Springer, Dordrecht.
Gillson, L. (2015). Biodiversity Conservation and Environmental Change: Using
Palaeoecology to Manage Dynamic Landscapes in the Anthropocene. Oxford University
Press, Oxford.
Hammond, W.M., Stone, M.E.B., Stone, P.A. (2020). Picture worth a thousand words:
Updating repeat photography for 21st century ecologists. Ecol. Evol., 10(1),
14113–14121.
Herrmann, S. and Tappan, G.G. (2013). Vegetation impoverishment despite greening: A case
study from central Senegal. J. Arid Environ., 90, 55–66.
Hoffman, M.T. and Cowling, R.F. (1990). Vegetation change in the semi-arid eastern Karoo
over the last 200 years: An expanding karoo-fact or fiction? Sth. Afr. J. Sci., 86, 286–294.
Hoffman, M.T. and O’Connor, T. (1999). Vegetation change over 40 years in the
Weenen/Muden area, KwaZulu-Natal: Evidence from photo-panoramas. Afr. J. Range
For. Sci., 16, 71–88.
Hoffman, M.T. and Rohde, R.F. (2007). From pastoralism to tourism: The historical impact of
changing land use practices in Namaqualand. J. Arid Environ., 70, 641–658.
Hoffman, M.T. and Rohde, R.F. (2011). Rivers through time: Historical changes in the
riparian vegetation of the semi-arid, winter rainfall region of South Africa in response to
climate and land use. J. Hist. Biol., 44, 59–80.
Hoffman, M.T. and Todd, S.W. (2010). Using fixed-point photography, field surveys and GIS
to monitor environmental change in Riemvasmaak, South Africa. In Repeat Photography:
Methods and Applications in the Natural Sciences, Webb, R., Boyer, D., Turner, R. (eds).
Island Press, Washington DC.
Hoffman, M.T., Rohde, R.F., Duncan, J., Kaleme, P. (2010). Repeat photography, climate
change and the long-term population dynamics of tree aloes in southern Africa. In Repeat
Photography: Methods and Applications in the Natural Sciences, Webb, R., Boyer, D.,
Turner, R. (eds). Island Press, Washington DC.
Hoffman, M.T., Rohde, R.F., Gillson, L. (2019). Rethinking catastrophe? Historical
trajectories and modelled future vegetation change in southern Africa. Anthropocene, 25,
100189.
Hoffman, M.T., Rohde, R.F., Gillson, L. (2020). Further comments on analysing trajectories
of vegetation and landscape change in southern Africa from historical field photographs.
Anthropocene, 32, 100274.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
68 Historical Ecology

Jacob, M., Frankl, A., Beeckman, H., Mesfin, G., Hendrickx, M., Guyassa, E., Nyssen, J.
(2015). North Ethiopian Afro-Alpine tree line dynamics and forest-cover change since the
early 20th century. Land Degrad. Dev., 26, 654–664.
Jacob, M., Frankl, A., Hurni, H., Lanckriet, S., De Ridder, M., Guyassa, E., Beeckman, H.,
Nyssen, J. (2017). Land cover dynamics in the Simien Mountains (Ethiopia), half a
century after establishment of the National Park. Reg. Environ. Change, 17(3), 777–787.
Kull, C.A. (2005). Historical landscape repeat photography as a tool for land use change
research. Nor. Geogr. Tidsskr., 59, 253–268.
Lanckriet, S., Rucina, S., Frankl, A., Ritler, A., Gelorini, V., Nyssen, J. (2015). Nonlinear
vegetation cover changes in the north Ethiopian highlands: Evidence from the Lake
Ashenge closed basin. Sci. Total Environ., 536, 996–1006.
Lätt, L. (2004). Eritrea re-photographed. Landscape changes in the Eritrean highlands
1890–2004. Thesis, University of Bern, Switzerland.
Masubelele, M., Hoffman, M.T., Bond, W. (2015). Biome stability and long-term vegetation
change in the semi-arid, south-eastern interior of South Africa: A synthesis of repeat
photo-monitoring studies. S. Afr. J. Bot., 101, 139–147.
McCain, C.T., Szewczyk, T., Knight, K.B. (2016). Population variability complicates the
accurate detection of climate change responses. Glob. Change Biol., 22, 2081–2093.
Meire, E., Frankl, A., De Wulf, A., Haile, M., Deckers, J., Nyssen, J. (2013). Land use and
cover dynamics in Africa since the nineteenth century: Warped terrestrial photographs of
north Ethiopia. Reg. Environ. Change, 13, 717–737.
Nyssen, J., Poesen, J., Haregeweyn, N., Parsons, T. (2008). Environmental change,
geomorphic processes and land degradation in tropical highlands. Catena, 75, 1–4.
Nyssen, J., Frankl, A., Haile, M., Hurni, H., Descheemaeker, K., Crummey, D., Ritler, A.,
Portner, B., Nievergelt, B., Moeyersons, J. (2014). Environmental conditions and human
drivers for changes to north Ethiopian mountain landscapes over 145 years. Sci. Total
Environ., 485, 164–179.
Nyssen, J., Frankl, A., Zenebe, A., Deckers, J., Poesen, J. (2015). Land management in the
northern Ethiopian highlands: Local and global perspectives, past, present and future.
Land Degrad. Dev., 26, 759–764.
O’Connor, T.G., Puttick, J.R., Hoffman, M.T. (2014). Bush encroachment in southern Africa:
Changes and causes. Afr. J. Range For. Sci., 31, 67–88.
Okubamichael, D.Y., Jack, S., Bösenberg, J.D.W., Hoffman, M.T., Donaldson, J.S. (2016).
Repeat photography confirms alarming decline in South African cycads. Biodivers.
Conserv., 25, 2153–2170.
Poulsen, Z. and Hoffman, M.T. (2015). Changes in the distribution of indigenous forest in
Table Mountain National Park during the 20th century. S. Afr. J. Bot., 101, 49–56.
Rogers, G.F., Malde, H.E., Turner, R.M. (1984). Bibliography of Repeat Photography for
Evaluating Landscape Change. University of Utah Press, Salt Lake City.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Use of Repeat Photography in African Historical Ecology 69

Rohde, R.F. and Hilhorst, T. (2001). After the fall: Political ecology and environmental
change in the lake manyara basin, Tanzania. Report, IIED Drylands Programme,
London.
Rohde, R.F. and Hoffman, M.T. (2012). The historical ecology of Namibian rangelands:
Vegetation change since 1876 in response to local and global drivers. Sci. Total Environ.,
416, 276–288.
Rohde, R.F., Hoffman, M.T., Durbach, I., Venter, Z., Jack, S. (2019). Vegetation and climate
change in the Pro-Namib and Namib desert based on repeat photography: Insights into
climate trends. J. Arid Environ., 165, 119–131.
Russell, J. and Ward, D. (2014). Vegetation change in northern KwaZulu-Natal since the
Anglo-Zulu war of 1879: Local or global drivers? Afr. J. Range For. Sci., 31, 89–105.
Scott, S.L., Rohde, R.F., Hoffman, M.T. (2018). Repeat landscape photography, historical
ecology and the wonder of digital archives in southern Africa. Afr. Res. Doc., 131, 35–47.
Shantz, H.L. and Turner, B.L. (1958). Photographic Documentation of Vegetational Changes
in Africa Over a Third of a Century. University of Arizona Press, Tucson.
Stuble, K.L., Bewick, S., Fisher, M., Forister, M.L., Harrison, S.P., Shapiro, A.M., Latimer,
A.M., Fox, L.R. (2020). The promise and perils of resurveying to understand global
change impacts. Ecol. Monogr., Article e01435, 1–14.
Turner, R.M., Ochung, H.A., Turner, J.B. (1998). Kenya’s Changing Landscape. University
of Arizona Press, Tucson.
Turner, R.M., Webb, R.H., Bowers, J.E., Hastings, J.E. (2003). The Changing Mile Revisited:
An Ecological Study of Vegetation Change with Time in the Lower Mile of an Arid and
Semiarid Region. University of Arizona Press, Tucson.
Vanha-Majamaa, I., Salemaa, M., Tuominen, S., Mikkola, K. (2000). Digitized photographs
in vegetation analysis – A comparison of cover estimates. Appl. Veg. Sci., 3, 89–94.
Venter, Z., Scott, S.L., Desmet, P.G., Hoffman, M.T. (2020). Application of Landsat-derived
vegetation trends over South Africa: Potential for monitoring land degradation and
restoration. Ecol. Indic., 113, 106206.
Von Hellerman, P. (2020). Partial stories: Repeat photography, narratives and environmental
change in Tanzania. Vis. Anthropol., 33, 363–391.
Ward, D., Hoffman, M.T., Collocott, S.J. (2014). A century of woody plant encroachment in
the dry Kimberley savanna of South Africa. Afr. J. Range For. Sci., 31, 107–121.
Webb, R.H., Boyer, D.E., Turner, R.M. (eds) (2010). Repeat Photography: Methods and
Applications in the Natural Sciences. Island Press, Washington DC.
Western, D. (2010). People, elephants, and habitat: Detecting a century of change using repeat
photography. In Repeat Photography: Methods and Applications in the Natural Sciences,
Webb, R., Boyer, D., Turner, R., (eds). Island Press, Washington DC.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Collapse of an iconic conifer: Long-term changes in the demography of Widdringtonia
White, J.D.M., Jack, S.L., Hoffman, M.T., Puttick, J., Bonora, D., February, E.C. (2016).

cedarbergensis using repeat photography. BMC Ecol., 16, 53.


Historical Ecology
70
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
6

Remote Sensing for


Historical Ecology
Pierre-Alexis HERRAULT1 and David SHEEREN2
1
University of Strasbourg, France
2
University of Toulouse, France

6.1. Introduction

Historical ecology involves reconstructing past landscapes and/or their


spatio-temporal trajectories to assess the effects of landscape history on current
biodiversity patterns (Jackson and Sax 2010). Given the wide range of species
response to disturbance, both spatially and temporally, remote sensing offers many
advantages for historical ecology through the vast amount of available spatial data
and multiple processing approaches.

This chapter provides an inventory of the available remote sensing approaches


for historical ecology at various spatio-temporal scales in relation to the existing and
accessible spatial data. This review is not intended to be exhaustive but offers
readers a glimpse of the possibilities and the stage of development of these
approaches. Two aspects will be highlighted throughout this chapter: (1) the
extraction of past landscapes from various available spatial sources and (2) the
production of spatio-temporal variables to model the effects of landscape history on
current biodiversity.

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
72 Historical Ecology

6.2. Landscape spatio-temporal changes as a proxy of biodiversity

Present-day biodiversity often reflects past land-use or past climate because of a


possible delay in the response of certain species to habitat disturbance (Kuussaari
et al. 2009). This interval depends on the changes themselves (i.e. intensity and
cyclicity) as well as species traits, namely their dispersal abilities. To assess the time
lag between changes in a landscape and changes in the associated populations,
studies usually rely on past and current spatial data that help highlight
spatio-temporal trajectories over large spatial extents and long periods.

In landscape ecology, biodiversity is most often studied from the perspective of


ecological habitats. Several fundamental ecological theories (e.g. island
biogeography, species-energy, species-heterogeneity) describe the relationships
between habitats and biodiversity, and these theories serve as a framework for
historical ecology. Therefore, a crucial step consists of describing specific habitat
properties by using adapted metrics. These properties can be distinguished as
follows:
– structural (or morphological) properties, which characterize the composition
(nature) and configuration (morphology, spatial organization) of a habitat;
– phyletic (or inherited) properties, which characterize the links between patches
of homologous habitats at different time slices;
– functional (or spectral) properties, which characterize the behavior of habitats
(productivity, phenology, etc.) that can be derived from high temporal resolution
signals that are recorded by satellite- and aircraft-based sensors.

The first two categories are available through the mapping of past landscapes at
different times (landscape states). Identifying the structure of a habitat requires
landscape metrics that permit characterizing habitat morphology (patch scale) or
spatial organization (patch-neighborhood or landscape scale). The phyletic
characteristics of a habitat and its evolution are obtained by linking homologous
patches across different time slices. Combining structural and phyletic properties
will ultimately provide evidence of past processes, such as the fragmentation or
simplification of landscapes. Finally, functional properties may be analyzed through
intra-annual or interannual biophysical variables that are calculated at a high
temporal frequency, as these variables are proxies for monitoring habitats and
habitat function over a relatively short period (seasonal cycles). Consequently, high
temporal resolution data are often required for measuring functional properties and
are, by default, available for more recent periods.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Remote Sensing for Historical Ecology 73

6.3. Mapping landscapes at different dates

Prior to describing the characteristics of ecological habitats, habitat patches must


first be extracted from the available spatial data for various time slices (Figure 6.1).
This initial step can be tedious depending on the quality (geometric, semantic,
temporal, spectral) of the available sources, thereby limiting the reliability of results
over extended areas. Here, we present various existing automated methods that
facilitate the extraction of landscape patches from commonly used spatial data or
that are potentially useful in historical ecology.

6.3.1. Airborne laser scanning data

Given the ability of LiIDAR to penetrate the forest canopy and detect
micro-features, LiDAR-derived data (Digital Terrain Model, DTM) can be used to
identify past ecosystems or provide additional information for previously discovered
sites (Dupouey et al. 2002). Because of the limited properties of these data, for
example, gray levels and variations in lighting, this step is usually undertaken
manually for historical ecology studies. This limitation is also related to a lack of
collaboration between historical ecology and archaeology, a field in which LiDAR
data have been exploited by researchers for many years.

Several approaches for detecting relatively old ecological habitats have been
proposed (Lambers et al. 2019), such as template-matching or knowledge-based
algorithms as well as object-based approaches or machine-learning methods. Of
these approaches, machine learning provides the best performance and consists of
calibrating the detection models using a set of learning samples having features
calculated from the digital terrain model. The use of decision trees, for example,
random forest, is generally successful; however, the results can be limited when
transposing the model outputs to another shape or to a different region. Current deep
learning methods show promise for the automatic mapping of archaeological relics
identified from LiDAR-based digital terrain models (Trier et al. 2019). Nonetheless,
the calibration of these neural networks requires an extensive set of learning
examples, an aspect that is not evident owing to the rare nature of archaeological
objects. The available image datasets (e.g. AlexNet, ImageNet) from which the
networks can be pretrained also contain insufficient examples for reproducing the
full variability of the forms assumed by these archaeological objects.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
74 Historical Ecology

Figure 6.1. Non-exhaustive overview of some of the major spatial sources available
for landscape/historical ecology studies at the landscape scale. RGB indicates the
red-green-blue spectral domain. n.s. indicates a “non-significant” temporal resolution
because of the non-systematic acquisition mode of the target mission. For a color
version of this figure, see www.iste.co.uk/decocq/ecology.zip

6.3.2. Historical maps

Old maps are regularly used in historical ecology (Cousins et al. 2015). They
provide a snapshot of landscapes from the 17th to 19th centuries and are commonly
the earliest time slice that is integrated into spatio-temporal landscape datasets. The
traditional approach of capturing the cartographic objects found on old maps relies
on user intervention (for digitizing). For numerous years, automated data-capture
techniques were developed to try and establish reproducible procedures. Most of the
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Remote Sensing for Historical Ecology 75

developed approaches, however, were specific to only one particular map and were
generally not applicable to other historical maps.

Recent work appears to be converging towards a standardization of the


processing chains (Herrault et al. 2013; Auffret et al. 2017). First, an initial cleaning
step is required systematically to eliminate elements that overlap the sought-after
data on the maps, for example, hatching, topographic contours, text. Smoothing
algorithms based on convolution or mathematical morphology are generally used.
Second, a classification step (object or pixel) is applied to assign each pixel or object
to a land-use class using supervised algorithms. More recently, neural networks
(deep learning) based on convolution kernels have been proposed for this step (Chen
et al. 2021). The obtained results show promise and a higher quality than those
obtained through traditional machine learning. Nonetheless, the parameterization of
these networks and the amount of required training data remain obstacles to their
wider application in historical ecology. Finally, a post-classification processing step
may be required, depending on the noise observed in the classification outputs. The
methods used for this processing are similar to those applied during the initial
cleaning step.

6.3.3. Old aerial photographs

Historical orthophotographs are primarily used to monitor landscape data from


the early 20th century (since the early 1930s) to the first systematic satellite
acquisitions in the early 1980s. These photographs provide information about
landscapes at a high spatial resolution (approximately 1 m) (Morgan et al. 2017) and
have a temporal resolution (time step between aerial surveys over the same
geographical area) of about 10 years. Despite these advantages, the delimitation of
habitats remains a tedious step because: (1) there is a high degree of heterogeneity
between surveys and even between images related to varying specifications and
quality (noise problem and vignetting); and (2) these data lack usable attributes,
particularly for panchromatic images, making automatic classification a very
complex task.

Approaches have been proposed recently to overcome these obstacles. First, a


spectralization/colorization process can be applied to enrich spectral attributes of the
images, particularly for B\&W photographs. The most effective approach at present
is based on deep-learning models, which have considerably improved the
colorization results (Goodfellow et al. 2014). The process involves training a neural
network using a dataset of paired panchromatic-color images to derive semantics
capable of generalizing the colorization to the whole target scene. In particular,
generative antagonistic networks (GAN) provide a much higher colorization quality
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
76 Historical Ecology

than that obtained with convolutional neural networks (CNN). Significant


improvements have also been demonstrated in the classification results obtained
from these newly colored images (Poterek et al. 2020). The production of old digital
surface models using photogrammetry techniques is also a promising avenue for
enriching attributes of old aerial missions and thus facilitate the extraction of
elements from these sources (ANR HIATUS, https://anr.fr/Projet-ANR-18-CE23-
0025).

Second, object-oriented approaches are most often favored for classifying aerial
photographs because they can process very high spatial resolution data. Several
studies have also demonstrated the utility of deep learning for the automatic
classification of aerial photographs (Sameen et al. 2018). Most rely on pretrained
CNN using available image datasets, such as AlexNet or ImageNet. These networks
have the advantage of teaching themselves the necessary features to classify at
different scales. They are therefore very effective for deriving contextual metrics
and often produce superior results to those obtained through traditional methods
based notably on textural attributes.

6.3.4. Satellite images

An exhaustive review of image satellite classification method is beyond the


scope of this chapter. Various recent publications provide an overview of existing
approaches (Dhingra and Kumar 2019). From a landscape ecology perspective, two
key aspects should be underlined:
– Classification outputs are strongly dependent on spatial, spectral and temporal
characteristics of the available satellite missions. These characteristics determine the
geometric and semantic quality of the produced land-cover maps.
– Machine-learning methods are currently the standard for image classification.
The quality of the learning examples and the selection of features are two critical
elements to their success. In recent years, the performance of classification
algorithms has rapidly improved through the use of CNN. The quality of the output,
however, remains highly dependent on the massive image datasets required for their
calibration. Although these datasets are well developed for the RGB domain, they
remain relatively poor in examples for the more resolved spectral domains, for
example, hyperspectral and radar data.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Remote Sensing for Historical Ecology 77

6.4. Modeling the effects of spatio-temporal changes on present-day


biodiversity

The ecological processes structuring biodiversity patterns can be disturbed by


forcing events of varying magnitude and intensity (Jackson and Sax 2010). These
changes are of several orders: (1) structural changes describing modifications in
landscape configuration and composition and (2) functional changes revealing
fluctuations in habitat properties, for example, productivity and phenology
(Figure 1.2).

6.4.1. Structural spatio-temporal metrics

Here, we present the existing structural spatio-temporal metrics according to two


major disturbances: landscape fragmentation and landscape homogenization.

6.4.1.1. Landscape fragmentation


The fragmentation of landscapes reduces habitat size and heightens habitat
isolation (Fahrig 2003). Assessing the effects on biodiversity follows two central
ecological theories: the theory of island biogeography, including species–area
relationships, and metapopulation theory.

The morphological description of habitat patches makes it possible to use


species–area relationships to assess biodiversity loss or gain in accordance with
patch attrition or expansion. This theory is also closely related to island
biogeography, which stipulates that at constant colonization rates, the extinction rate
of species is likely to decrease (increase) as the area increases (decreases) through a
positive effect of size on richness. Larger patches are generally more heterogeneous
than smaller ones and therefore favor a greater diversity of resources for different
species. Thus, the morphology of habitat patches can be characterized simply using
surface, perimeter or shape metrics. This implies the availability of geometric
objects (most often polygonal) at different times for one or more habitat types. At
the landscape scale, the colonization/extinction dynamics of species also depend on
the connectivity of the habitat patches. For a given surface area, the colonization rate
of a habitat patch is lower when the patch is further away from the source; the
diversity will therefore be higher for a less isolated habitat. The degree of
complexity of the connectivity metrics is variable (Kindlmann and Burel 2008), and
their use most often depends on the specific research objectives and the semantic
quality of land-cover maps, i.e. the information available through the landscape
matrix.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
78 Historical Ecology

The effects of habitat disturbances on biodiversity are most often modeled by


comparing the characteristics of past and present landscapes. The difference
between current richness and the value predicted by the statistical model on the basis
of past landscapes provides an estimate of the post-disturbance species debt or
credit. This type of modeling has been widely adopted for different taxa (Figueiredo
et al. 2019) but may be questioned because of the possibilities of over- or
underestimating extinction processes. Other approaches aim to integrate the
temporal trajectory of habitats into species-habitat models by considering habitat
changes as a cumulative process. Consequently, in addition to the present habitat
characteristics, accounting for variables such as the rate of area change and
connectivity may considerably improve the explanation of species richness and
abundance (Herrault et al. 2016). Finally, the life history of habitat patches (their
phyletic properties; Figure 1.2) over time appears to be a relevant variable for
explaining current biodiversity patterns (Ewers et al. 2013). The historical separation
of habitat fragments is represented by a landscape “terrageny” in a similar manner as
phylogeny represents the evolutionary divergence among species.

6.4.1.2. Landscape homogenization


The homogenization of a landscape reduces its capacity to support a large
number of species owing to the lack of complementary resources to meet the needs
of a large pool of species. This results in a decline in overall diversity (taxonomic
diversity) as well as functional diversity, with specialists being replaced gradually
by generalist species (Gámez-Virués et al. 2015).

Landscape homogenization is characterized both as a decrease in the diversity of


land-use (heterogeneity of composition) and as an increase of habitat patch size
(heterogeneity of configuration; Fahrig et al. 2011) The configurational
heterogeneity is primarily based on morphological metrics (area, length) and
summarized (summed or averaged) at the scale of the neighborhood of the habitat
patches. The compositional heterogeneity is measured using indicators of land-use
diversity, such as the relative share of land-use in the neighborhood of the patch
and/or Shannon metrics, which have the benefit of being less correlated with
configuration metrics. Modeling the effects of homogenization on biodiversity is
most often undertaken via statistical models relating the characteristics of the current
landscape to current biodiversity. Other researchers introduce temporal composition
metrics (absolute or relative, for example, ΔShannon, ΔWood) directly into
ecological distribution models (Bonthoux et al. 2013). It should be noted that this
latter approach remains marginal among the cited publications, notably because of
the difficulties in reconstructing past land-use trajectories. Moreover, these metrics
imply having a detailed and uniform land-use nomenclature for the considered time
trajectory (at different times). This can be a limiting factor for older periods where
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Remote Sensing for Historical Ecology 79

the available sources are not always of sufficient semantic quality relative to the
established nomenclature criteria. Image classification errors can also limit the
reliability of derived spatio-temporal trajectories by leading to a loss of information
at each time slice. Finally, the strict delimitation of objects in the landscape does not
always reveal the details and continuity of a landscape.

Given these latter issues, a continuous approach (including no classification step)


has been proposed and contrasts with the above-mentioned discrete vision. These
works are based on the spectral variation hypothesis (SVH), which assumes that the
spectral dispersion in the image (NDVI, texture or reflectance values) is correlated
with the heterogeneity of the habitat patches in the landscape. Many approaches
have been proposed to quantify spectral heterogeneity and relate it to alpha and beta
diversities (Rocchini et al. 2010). They include (1) dispersion-based metrics, which
rely on calculations of the standard deviation or the coefficient of variation of pixel
values; (2) spectral distance-based metrics, which involve an initial classification of
pixels with non-supervised methods, such as K-means methods, to describe
similarities between spectral classes. Landscape heterogeneity is then calculated
with diversity indices (e.g. Shannon, Rao; Rocchini et al. 2021). Results suggest that
spectroheterogeneity is an effective and quickly produced proxy of biodiversity;
however, the quality of its output is strongly dependent on (1) the characteristics of
satellite or airborne sensors; (2) scale-matching problems between the remotely
sensed and field-based diversity data, and (3) the types of species diversity
measures.

6.4.2. Functional spatio-temporal metrics

Functional metrics require high temporal resolution spatial data to capture the
behavior of ecological habitats (Figure 6.2). These spectral–temporal metrics refer to
different aspects of flora and fauna habitat such as productivity, phenology, moisture
or snow cover.

Productivity metrics are most often obtained through calculating biophysical


variables from various satellite images. Nonetheless, few satellite missions provide
sufficient temporal characteristics, for example, frequency and age, to highlight the
seasonal behavior of landscapes and their past interannual variations. The MODIS
mission, for example, provides high-frequency NDVI data capturing the vegetation
photosynthetic activity considered as a proxy of primary productivity. Composite
products have therefore been proposed, for example, dynamic habitat indices (DHI),
and make it possible to simultaneously assess the role of different factors (vegetation
stress (DHI_min), seasonal variations (DHI_var) and total productivity (DHI_cum))
on bird, mammal and amphibian diversity patterns (Radeloff et al. 2019). Few
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
80 Historical Ecology

studies, however, have tested DHI interannual trends as a potential explanatory


variable of current biodiversity patterns. One reason could be the low spatial
resolution of MODIS that may hinder its use owing to a spatial mismatch with field
data. Despite a lower temporal resolution (16 days), Landsat data are also relevant
for deriving functional productivity metrics. The succession of Landsat sensors
(since the early 1980s) and the recent Sentinel 2 missions make a major asset for
reconstructing the past annual productivity of habitat patches over a longer period of
time.

Figure 6.2. Landscape spatio-temporal reconstruction based on the


properties of ecological habitats. For a color version of this
figure, see www.iste.co.uk/decocq/ecology.zip

Spatial and temporal variations in phenology are critical factors in resource


availability and habitat use for a wide range of species. Changes in phenological
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Remote Sensing for Historical Ecology 81

timing can lead to dysfunctions in the interactions between species and, more
specifically, to cascading asynchronies across trophic levels (Mayor et al. 2017).
Such indices are based on the initial calculation of seasonal indices, such as the day
of the start/end of season, peak greening. They can be derived from the intra-annual
curves of vegetation indices, for example, EVI and NDVI. Consequently, high
temporal resolution satellite data are required to capture these seasonal metrics. For
instance, the MODIS mission provides a phenological product (MCD12Q2) having
different bands that display these transition dates. The Landsat and Sentinel 2
missions are also well adapted to derive these metrics. Nonetheless, the moderate
temporal resolution and low temporal depth remain two limiting factors for
reconstructing spatio-temporal phenological trajectories. Developing synergies
between these two missions will be critical for future studies in historical ecology.

6.5. References

Auffret, A.G., Kimberley, A., Plue, J., Skånes, H., Jakobsson, S., Waldén, E.,Wennbom, M.,
Wood, H., Bullock, J.M., Cousins, S.A. et al. (2017). Histmapr: Rapid digitization of
historical land-use maps in r. Meth. Eco. Evo., 8(11), 1453–1457.
Bonthoux, S., Barnagaud, J.-Y., Goulard, M., Balent, G. (2013). Contrasting spatial and
temporal responses of bird communities to landscape changes. Oecologia, 172(2),
563–574.
Chen, Y., Carlinet, E., Chazalon, J., Mallet, C., Duménieu, B., Perret, J. (2021). Combining
deep learning and mathematical morphology for historical map segmentation, arXiv
preprint arXiv:2101.02144.
Cousins, S.A., Auffret, A.G., Lindgren, J., Tränk, L. (2015). Regional-scale land-cover
change during the 20th century and its consequences for biodiversity. AMBIO A Journal
of the Human Environment, 44(1), 17–27.
Dhingra, S. and Kumar, D. (2019). A review of remotely sensed satellite image classification.
Int. Jour. Elec. Comp. Eng. (2088–8708), 9(3).
Dupouey, J.-L., Dambrine, E., Laffite, J.-D., Moares, C. (2002). Irreversible impact of past
land use on forest soils and biodiversity. Ecology, 83(11), 2978–2984.
Ewers, R.M., Didham, R.K., Pearse, W.D., Lefebvre, V., Rosa, I.M., Carreiras, J.M., Lucas,
R.M., Reuman, D.C. (2013). Using landscape history to predict biodiversity patterns in
fragmented landscapes. Eco. Lett., 16(10), 1221–1233.
Fahrig, L. (2003). Effects of habitat fragmentation on biodiversity. Ann. Rev. Eco. Evo. Sys.,
34(1), 487–515.
Fahrig, L., Baudry, J., Brotons, L., Burel, F.G., Crist, T.O., Fuller, R.J., Sirami, C.,
Siriwardena, G.M., Martin, J.-L. (2011). Functional landscape heterogeneity and animal
biodiversity in agricultural landscapes. Eco. Lett., 14(2), 101–112.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
82 Historical Ecology

Figueiredo, L., Krauss, J., Steffan-Dewenter, I., Sarmento Cabral, J. (2019). Understanding
extinction debts: Spatio-temporal scales, mechanisms and a roadmap for future research.
Ecography, 42(12), 1973–1990.
Gámez-Virués, S., Perovic, D.J., Gossner, M.M., Börschig, C., Blüthgen, N., De Jong, H.,
Simons, N.K., Klein, A.-M., Krauss, J., Maier, G. et al. (2015). Landscape simplification
filters species traits and drives biotic homogenization. Nat. Com., 6(1), 1–8.
Goodfellow, I., Pouget-Abadie, J., Mirza, M., Xu, B., Warde-Farley, D., Ozair, S., Courville,
A., Bengio, Y. (2014), Generative adversarial networks. Advances in Neural Information
Processing Systems. ACM, New York, 26722680.
Herrault, P.-A., Sheeren, D., Fauvel, M., Paegelow, M. (2013). Automatic extraction of
forests from historical maps based on unsupervised classification in the CIELab color
space. In Geographic Information Science at the Heart of Europe, Vandenbroucke, D.,
Bucher, B., Crompvoets, J. (eds). Springer, Wiesbaden.
Herrault, P.-A., Larrieu, L., Cordier, S., Gimmi, U., Lachat, T., Ouin, A., Sarthou, J.-P.,
Sheeren, D. (2016). Combined effects of area, connectivity, history and structural
heterogeneity of woodlands on the species richness of hoverflies (diptera:Syrphidae).
Land. Eco., 31(4), 877–893.
Jackson, S.T. and Sax, D.F. (2010). Balancing biodiversity in a changing environment:
Extinction debt, immigration credit and species turnover. Tren. Eco. Evo., 25(3),
153–160.
Kindlmann, P. and Burel, F. (2008). Connectivity measures: A review. Land. Eco., 23(8).
879–890.
Kuussaari, M., Bommarco, R., Heikkinen, R.K., Helm, A., Krauss, J., Lindborg, R., Öckinger,
E., Pärtel, M., Pino, J., Rodà, F. et al. (2009). Extinction debt: A challenge for
biodiversity conservation. Tren. Eco. Evo., 24(10), 564–571.
Lambers, K., Verschoof-van der Vaart, W.B., Bourgeois, Q.P. (2019). Integrating remote
sensing, machine learning, and citizen science in Dutch archaeological prospection. Rem.
Sen., 11(7). 794.
Mayor, S.J., Guralnick, R.P., Tingley, M.W., Otegui, J., Withey, J.C., Elmendorf, S.C.,
Andrew, M.E., Leyk, S., Pearse, I.S., Schneider, D.C. (2017). Increasing phenological
asynchrony between spring green-up and arrival of migratory birds. Sci. Rep., 7(1), 1–10.
Morgan, J.L., Gergel, S.E., Ankerson, C., Tomscha, S.A., Sutherland, I.J. (2017). Historical
aerial photography for landscape analysis. In Learning Landscape Ecology, Gergel, S.E.
and Tuner, M.G. (eds). Springer, Cham.
Poterek, Q., Herrault, P.-A., Skupinski, G., Sheeren, D. (2020). Deep learning for automatic
colorization of legacy grayscale aerial photographs. IEEE Jour. Sel. Top. App. Ear. Obs.
Rem. Sen., 13, 2899–2915.
Radeloff, V., Dubinin, M., Coops, N., Allen, A., Brooks, T., Clayton, M., Costa, G.,Graham,
C., Helmers, D., Ives, A. et al. (2019). The dynamic habitat indices (DHIs) from modis
and global biodiversity. Rem. Sen. Env., 222, 204–214.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Remote Sensing for Historical Ecology 83

Rocchini, D., Balkenhol, N., Carter, G.A., Foody, G.M., Gillespie, T.W., He, K.S., Kark, S.,
Levin, N., Lucas, K., Luoto, M. et al. (2010). Remotely sensed spectral heterogeneity as a
proxy of species diversity: Recent advances and open challenges. Eco. Inf., 5(5),
318–329.
Rocchini, D., Salvatori, N., Beierkuhnlein, C., Chiarucci, A., de Boissieu, F., Förster, M.,
Garzon-Lopez, C.X., Gillespie, T.W., Hauffe, H.C., He, K.S. et al. (2021). From local
spectral species to global spectral communities: A benchmark for ecosystem diversity
estimate by remote sensing. Eco. Inf., 61, 101195.
Sameen, M.I., Pradhan, B., Aziz, O.S. (2018). Classification of very high resolution aerial
photos using spectral-spatial convolutional neural networks. Jour. Sens., 2018, Article ID
7195432, 1–12.
Trier, Ø.D., Cowley, D.C., Waldeland, A.U. (2019). Using deep neural networks on airborne
laser scanning data: Results from a case study of semi-automatic mapping of
archaeological topography on arran, Scotland. Arch. Pros., 26(2), 165–175.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
7

Soil Archives: Where


Soilscape History Meets
Present-day Ecosystems
Boris BRASSEUR1, Damien ERTLEN2 and Vincent ROBIN3
1
Jules Verne University of Picardie, Amiens, France
2
University of Strasbourg, France
3
University of Lorraine, Metz, France

7.1. Introduction

The interpretation of soil archives, although complex, is nevertheless essential


for understanding the impact of soils on present-day environments and deciphering
the environmental history of a site. Soil archives are local, although within a
continuous soilscape, they allow the study of spatial structures. Compared with
sedimentary archives, they do not strictly follow stratigraphic principles. Combining
superficial microtopography, buried biological remains, bioturbation and chemical
evolution, these archives appear as mixed pieces of old and new disassembled
jigsaw puzzles. Such puzzles are snapshots of past environments, and understanding
the biological and physicochemical pedogenic processes is the main way to
disentangle them. This chapter presents some key elements for understanding the
archiving process through to the reconstruction of a fragment of paleoenvironmental
history. The presented key elements are associated with our respective experiences,
and we do not try to present an exhaustive list.

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
86 Historical Ecology

7.2. Mechanisms of soil archiving and the associated dynamics

7.2.1. Pedoturbations of biological and physical origins

When dealing with soil archives, a key issue to consider is that soils are dynamic
entities. Biological and/or physical processes induce the mixing of the soil matrix
down to a fine spatial scale (i.e. micromorphological scale). Indeed, when
environmental conditions permit biological activities in soil, the related bioturbation
causes a vertical transfer of elements. These transfers are mainly related to the
presence of micropassages in which soil moves downward when they collapse.
These spaces result from the activity of small mammals (e.g. Talpidae) and
earthworms (e.g. Lumbricidae). The latter of these also move worm casts – matter
excreted by the worms – upward to the soil’s surface (Lavelle et al. 2016). The
mixing of soil also occurs through physical processes. These processes include tree
uprooting, in which soil embedded in the root plate (Šamonil et al. 2010) moves
upward through the overturning of the tree trunk and roots. Cryoturbation stems
from the freeze–thaw-induced movement of the entire active layer portion of the soil
(Stinchcomb et al. 2014). Argillipedoturbation (swelling clay pedoturbation) and
seismiturbation (earthquake pedoturbation) are also important physical soil
disturbances. Thus, the soil matrix or soil aggregates may move in any direction,
depending on the type and intensity of soil biological activities, the on-site
vegetation cover, local climate, seasonality, soil clay content and local tectonic
activity. These processes sometimes preserve the soil’s environmental memory, but
they also represent a substantial challenge for researchers trying to read the soil
archive.

7.2.2. Eluviation–Illuviation

The percolation of water through the soil is an important process in humid


climates. This downward movement of water transports fine particles (e.g. clays,
fine silt, organic matter) and molecules (e.g. organic matter, oxyhydroxides,
carbonates) capable of passing through the pore spaces of the soil matrix. These two
components are transported, respectively, in suspension or dissolved in the water as
they pass through the pore space. A soil layer subjected mainly to the departure of
particles is termed “eluvial” (E horizons) and often undergoes partial discoloration
(Figure 7.1). With the downward change in the texture of the horizons (i.e. clay
accumulation), the disappearance of larger pores (thus a reduced movement through
the soil), the pH gradient of the soil and other physical obstacles (i.e. blocks of
limestone), the eluviated soil particles are likely to flocculate and be adsorbed at
depth, and the molecules will also precipitate at a deeper soil horizon. The soil area
receiving the material transported by gravity water is termed “illuvial” (B horizons).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Soil Archives 87

These eluviation–illuviation dynamics resulting from precipitation will, over the


centuries, determine the structure and texture of soils. With these
transfers/translocations of particles and molecules, many parameters fundamental to
plant growth (water storage capacity of soil horizons, volume density, CEC,
resistance to penetration) are influenced over these longer time scales.

Figure 7.1. Soil profile showing distinct horizons (labeled on the right-hand side) that
mark the eluviation and illuviation of organic matter (note the dark horizontal stripes
at approximately 50 cm depth) and metal oxides (approximately 65–85 cm depth) at
the expense of the initial parent material. This type of soil, a podzol, is dominated by
these vertical redistributions and little bioturbation. For a color version of this figure,
see www.iste.co.uk/decocq/ecology.zip

Human influence through land-use (field, pasture, forest) can be very significant
and long-lasting. For example, human modification of pH can affect
eluviation–illuviation patterns. In the case of eluviation–illuviation of clays, Ca2+
and Al3+ ions are flocculants and will prevent clay movement where these ions are
abundantly available in the soil solution. The eluviation–illuviation of clays is
therefore at a maximum under pH conditions of moderate acidity (5.5–6.5), an
acidity level associated with minimum Ca2+ and Al3+ concentrations in soil water.
This soil acidity is a parameter that has been controlled by humans through
agriculture for several millennia (see secton 7.2.1), and thus, human activities alter
eluviation–illuviation patterns and control the progressive vertical organization of
fine particles, soluble molecules and ions in the soil. These vertical redistributions
are in an unstable dynamic equilibrium with natural biological mixing, which
disturbs the soil’s vertical organization.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
88 Historical Ecology

7.2.3. Anthropogenic factors

In an anthropic landscape, various direct soil disturbances will lead to the


archiving of episodes from the local land-use history. A common case is that of pits
or depressions dug in the soil. Once the continual dredging/maintenance of these
depressions is stopped, these features will accumulate sedimentary particles until
filled. These particles, subjected to local natural and anthropogenic dynamics,
originate mainly from local superficial horizons and may include plant and animal
remains, as well as cultural artifacts. Such depressions can sometimes bear witness
to now-truncated soils and often have unique properties (grain size texture,
chemistry, structure) that affect aboveground communities.

A major anthropogenic disturbance of the world’s cultivable soils, although


superficial in terms of depth, concerns the impact of agricultural implements (hoes,
ploughs, etc.). The effect on the surface and the superficial horizons of the soils
leads to a leveling of micro- and mesotopographic landscapes and the disturbance of
the superficial horizons. This anthropogenic effect will erase the inherited
topography and reset the palimpsest of the microtopography of soils. Among the
cultivable lands in western Europe, it is very difficult, except for mountain areas, to
find areas that have not experienced past agricultural activity (Sabatini et al. 2018).
In contrast, in a forest context characterized by a dense and continuous vegetation
cover, the soil surface is protected/stabilized and therefore the conservation of its
landforms is better.

The concept of anthrosol (IUSS Working Group WRB 2015) includes all soils
that have been formed or deeply modified by human activities since the beginning of
the Holocene. Given the vast quantity of soils that have been affected directly or
indirectly by human activities, this definition is impractical for the pedologist.

7.2.4. Effects of geomorphological processes on soil archives

In the three previous sections, we described the biological, physical and


anthropological archiving elements of pedological processes. Geomorphological
processes also affect the soilscape and can produce excellent soil and sediment
archives or …destroy them. Aeolian, fluvial and gravitational processes are the
major erosional forces that affect a soil’s environmental memory. Soilscapes lacking
the protection of vegetation are very sensitive to ablation (Figure 7.2(b)).
Consequently, biomes with little vegetation and agrosystems with denudated soils
are not favorable for directly preserving soil archives.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Soil Archives 89

Tracing soil archives is difficult in the case of transport by aeolian processes or


by large fluvial systems because the constituents are mixed over a large area and the
spatio-temporal information is lost entirely. In a small catchment or at the slope
scale, soil material is often stored as pedosedimentary layers, and the properties of
eroded soils can be traced. Even if the soil architecture is lost, some chemical
parameters or organic botanical remains (e.g. seeds, charcoal, phytoliths) can be
preserved (Figure 7.2(d)).

The former soil properties can be even better preserved when the horizons are
buried under other layers of sediment (Figure 7.2(c)). In these cases, the soil
parameters or constituents can be considered as fossilized (Yaalon 1971). These
paleosols are no longer influenced by present-day climate, vegetation, farming, etc.,
and can be precisely dated. This is not the case of preserved soils that are in
equilibrium with present-day conditions. The surficial soils also contain inherited
properties and archives (see sections 7.2.1, 7.2.2 and 7.2.3), but these soils are often
difficult to date because they are the result of a temporal continuum (Figure 7.2(a)).

Figure 7.2. The effects of the main geomorphological processes


on the preservation and quality of soil archives. For a color
version of this figure, see www.iste.co.uk/decocq/ecology.zip

On the other hand, because they are still part of the present-day ecosystems,
these soils remain relevant for historical ecology studies.

When looking for archives in soils, the main error to avoid is confusing soil
horizons with soil strata. Both appear as a vertical succession of layers; however, the
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
90 Historical Ecology

strata are the result of successive deposits in time that are fossilized. They can
therefore be read as a relative chronology from the bottom (oldest) to the top
(youngest) of the sequence. This is not the case of soil horizons that are still
subjected to reorganization processes (see sections 7.1.1, 7.1.2 and 7.1.3).
Consequently, no relative chronology can be read from the soil horizons, and the age
or residence time of every incorporated constituent must be considered.

7.3. Examples of soil archives and their influence on current


ecosystems

7.3.1. Chemical archives, witnesses of progressive soil transformations

The chemical parameters of soils (i.e. pH, C, N, P, exchangeable cations) are


very rarely in equilibrium and are therefore subject to evolutionary dynamics at an
annual to millennial scale. The inherited initial state is the starting point for these
dynamics; for example, this could be a parent material newly accessible to
pedogenesis. It can also be a tipping point, such as climate or land-use change. For
the chemical parameters of soils, the subsequent evolutionary dynamics will tend
towards a new equilibrium state. The history of a soil is therefore important for
explaining the state of current chemical characteristics as well as for understanding
the successions of past and future states. The associated soil biotas are subject to the
influence of these chemical parameters and are likely to be structured by the
achieved evolutionary trajectory.

To illustrate this point, we can look at soil pH, one of the most structuring
parameters for edaphic ecosystems. Under humid climates, soil pH is subject to
acidification by the leaching and/or the input of protons (from atmospheric
precipitation, litter decomposition, root exudation and respiration) in the upper part
of the soil (Slessarev et al. 2016). This acidity increase reduces the quality and
quantity of the harvests of many food crops. To fight against this natural process of
acidification, farmers, especially in northwestern Europe, have added limestone
amendments for more than two millennia, thus neutralizing the acidity of the soil
(see Pliny the Elder: Book XVII-Chapter 4). A spatio-temporal study of silty soils in
northern France compared two chronosequences of land-use change (Figure 7.3).
The authors compared plots that had been deforested and cultivated within the past
few years to several centuries and plots where afforestation (on cultivated land)
occurred recently to nearly 1.6 millennia ago (Brasseur et al. 2018). The soil pH of
the deforested chronosequence revealed that the tilled surface horizon, initially
acidic, will require a few years to approach neutrality, whereas this neutralization
will take one to three centuries at more than 1-meter depth.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Soil Archives 91

Conversely, when cultivated land (with neutral pH) is abandoned and afforested,
the natural acidification dynamics are slower. Indeed, after 1.6 millennia of
afforestation of previously cultivated land, although the pH at the surface has
decreased, the pH of deeper horizons has lowered to a much lesser extent, indicating
a slow process that is still underway.

Figure 7.3. Mean (n = 6; ± SE) pH(water) profiles in luvisols from the loess plateau in
northern France characterized by differences in deforestation and afforestation ages.
Left-hand graph: pH(water) profiles for agricultural plots that were afforested at
various times in the past. Right-hand graph: plots deforested at various times in the
past for agricultural use. AF: ancient forest (circa. 1,600 years BP), AC: ancient
culture (>500 years BP), LUC: land-use change. For a color version of this figure, see
www.iste.co.uk/decocq/ecology.zip

The characterization of these evolutionary soil dynamics makes it possible to


identify the successions of ongoing and forthcoming chemical states for soils in
similar contexts (parental material, climate, land-use). This work also reveals that
the past agricultural use in the now-forested lands of northwest Europe can still be
characterized owing to this very gradual acidification of the soil. Lowland temperate
forests with a primeval character no longer exist in Europe (Peterken 1996), and
most European forests have been cultivated in a more or less distant past (Sabatini
et al. 2018). Thus, soil pH profiles may be an indicator of afforestation oldness in
many forested areas. A wealth of literature also describes the long-lasting inherited
chemical effects of past human agricultural activities on soils that can be deciphered
from the present-day plant community composition (e.g. Dambrine et al. 2007).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
92 Historical Ecology

7.3.2. Physical archives: reading the soil pit profile and


microtopographic features

An essential method for accessing the in situ physical properties of the soil
archives, for example, horizonation, soil structure and disturbance, and density
anomalies, is studying soil pits. This approach is the best way to describe the spatial
relationships between horizons, the soil and parent material, pedofeatures (e.g.
redoximorphic and textural characteristics, secondary precipitations), and biological
and physical pedoturbations. Reconstructing the history of the soil then consists of
unraveling the spatial relationships to derive a chronological sequence of events that
have affected the soil. We can do this using simple logical principles borrowed from
stratigraphy, including superposition (limited use, see sections 7.1.2 and 7.1.4),
inclusion and cross-cutting. The progression in horizon individualization from the
initial parent material is also a key criterium. As an example, the podzol profile
(Figure 7.1) evolves through the following steps: 1) the parent material is accessible
to pedogenesis, 2) acidification, horizonation and podsolization, 3) formation of
dark horizontal stripes and 4) intense reworking (cross-cutting) of topsoil and
eluviated horizons by the intrusion of roots and burrows following a possible local
nutrient enrichment. Such enrichment can reactivate the biological activity (and
bioturbation) of these very nutrient-poor soils (Robin 2005).

Figure 7.4. Aerial photography and DEM of the same area with ancient forested (1, 2, 3)
and deforested (4, 5; cleared ca. 150 years ago) sectors. 1) Track ways, 2) charcoal
kilns, 3) demarcation of old (probably Roman) agricultural plots, 4) less visible under
pasture and 5) no demarcation under culture. The darker spots are former charcoal
kilns. For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip

Since the 2000s, laser remote sensing techniques have revolutionized the
observation of soil microtopography. The generated digital elevation models (DEM)
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Soil Archives 93

offer the opportunity to observe and interpret landscapes invisible to the eye,
especially for land having a forest cover. In the Thiérache forest (near
franco-belgian frontier), for example, DEM (Figure 7.4) has revealed numerous old
agricultural parcel limits, ancient trackways and charcoal kilns. In recently
deforested areas, such memories of land-use disappear gradually under pastures and
much more quickly under the recurrent impact of ploughing in cultivated fields.

7.3.3. Soil organic matter

Soil organic matter (SOM) is a functional category of soil constituent. Its quality,
chemical structure, quantity and age vary within the soil profile and the succession
of horizons. Transport processes and biological and anthropological transformations
(see sections 7.1.1 and 7.1.2) control these distributions. Vegetation and the
microbial mass that transforms this vegetation are the main sources of OM
(Kögel-Knabner 2002). The OM derives mainly from the degradation of plant
fragments within the litter or from root exudates. The extraordinary biochemical
diversity of the plants is partly reflected by the SOM even after transformation and
microbial degradation. When researchers manage to trace and date SOM, it can
serve as a proxy of past vegetation and past landscape dynamics. SOM contains
thousands of different organic molecules. The classification of these complex
components depends on the tools at hand and the scale of the study.

Figure 7.5. Soil organic matter determined by near-infrared spectroscopy. (a)


Near-infrared spectra of soils and (b) multiple discriminant analysis on a set of 1,333
reference spectra from 75 sites under temperate forest and grassland cover. For a
color version of this figure, see www.iste.co.uk/decocq/ecology.zip
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
94 Historical Ecology

In the last 20 years, numerous studies using gas chromatography–mass


spectroscopy (GCMS) and other powerful methods have identified some very
specific molecules or isotopic ratios that identify particular categories of vegetation
or sometimes specific species (Trendel et al. 2010). Because they are
time-consuming, these methods are often applied to only a few soil profiles and are
of little use for landscape-scale studies. The fingerprinting approach of SOM by
near-infrared spectroscopy is less accurate and functions as a black box in regard to
SOM molecular chemistry, but it is fast and effective and provides an overall picture
of the SOM. As well, the existence of an extensive catalogue of reference soil
samples with a well-known ecological history (Ertlen et al. 2010; Figure 7.5) allows
near-infrared spectroscopy to discriminate SOM according to its plant-association
source. It is thus an effective tool for historical and landscape ecology.

7.3.4. Botanical remains

Soils are not well suited for the long-term preservation of organic botanical
remains. Biogeochemical activities induce the degradation of organic material,
limiting their retrospective representativeness. Such is the case of plant
macroremains that are assimilated into the carbon cycle by organic degradation.
Nonetheless, several botanical remains have specific features that make them
resistant to biogeochemical degradation. Once these botanical remains are extracted
from soil, taxonomic analyses of these samples can shed valuable insight into the
studied organisms and provide a portrait of the environmental context (Ferguson
2005). This resistance to degradation is observed with certain microfossils such as
pollen grains, owing to their exine composed of sporopollenin, or phytoliths, being
microscopic structures composed of silica. These microfossils, when found in
significant quantities in soils, can provide information about past land-use and
vegetation history (e.g. Robin et al. 2012; Contreras et al. 2014).

These indicators are often difficult to use, however, because establishing the
chronological framework is a challenge in a soil. Assessing the chronology of an
archive is a key step in any retrospective study. Indeed, most soils are continuously
mixed (see section 7.1.1); therefore, soils do not present a chronologically structured
stratification (Carcaillet 2001). This contrasts with chrono-stratified sedimentary
sequences, such as those recovered from lakes or mires, in which chronological
archiving is nearly continuous from the sequence bottom to the sediment surface.
Only in the case of pedosedimentary accumulation that have clear boundaries
(Figure 7.2(d)) is it possible to identify a relative stratigraphy in soil archives (Robin
et al. 2014). However, more precise chronological insights can be obtained from
charred botanical remains. Such indicators can be preserved for a long period and
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Soil Archives 95

are datable directly by 14C measurements (see section 7.1.4). The carbonization of
botanical remains, possibly related to anthropic or natural fires, transforms them into
inorganic carbon remains, allowing them to resist biogeochemical degradation.
Moreover, charcoal remains taxonomically identifiable, as its anatomical structures
are well preserved during carbonization, except for some deformation and shrinkage
(Braadbaart and Poole 2008). Quantitative and qualitative analyses of the charcoal
assemblages from soils provide insight into fire and forest vegetation history (Nelle
et al. 2013). Charcoal pieces must be large enough to be taxonomically identifiable
(i.e. megacharcoal, >800 µm). Therefore, soil megacharcoal assemblages are
relevant in fine-scale studies (Feiss et al. 2017) or at a catchment scale when, for
example, studying erosive events (Robin et al. 2014).

Overall, investigations that combine diverse types of botanical remains or


diverse types of indicators are the most relevant to obtain the best possible, reliable
and complete picture of past environments (Robin et al. 2012; Contreras et al. 2014).

7.4. Conclusion

Soils are a complex environmental interface where the chronological succession


of environments can produce distinct archives having a complex interpretation,
much like studying an environmental palimpsest. In this context, it is best to
combine the skills of different specialists in the (archaeo-) environmental sciences.
The combined study of the different soil archives (plant remains, SOM, chemical
parameters, macro- and micromorphological features, microtopography,
archaeology) allows an increasingly less lacunary picture to emerge of the
environmental history and its legacies in the soil; an environmental story whose
heritage, while fading, still structures today’s terrestrial ecosystems.

7.5. References

Braadbaart, F. and Poole, I. (2008). Morphological, chemical and physical changes during
charcoalification of wood and its relevance to archaeological contexts. J. Archeol. Sci.,
35(9), 2434–2445.
Brasseur, B., Spicher, F., Lenoir, J., Gallet-Moron, E., Buridant, J., Horen, H. (2018). What
deep-soil profiles can teach us on deep-time pH dynamics after land use change? Land
Degrad. Dev., 29(9), 2951–2961.
Carcaillet, C. (2001). Are Holocene wood-charcoal fragments stratified in alpine and
subalpine soils? Evidence from the Alps based on AMS 14C dates. Holocene, 11(2),
231–242.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
96 Historical Ecology

Contreras, D.A., Robin, V., Gonda, R., Hodara, R., Dal Corso, M., Makarewicz, C. (2014).
(Before and) after the flood: A multiproxy approach to past floodplain usage in the middle
Wadi el-Hasa, Jordan. J. Arid Environ., 110, 30–43.
Dambrine, E., Dupouey, J.-L., Laüt, L., Humbert, L., Thinon, M., Beaufils, T., Richard, H.
(2007). Present forest biodiversity patterns in France related to former roman agriculture.
Ecology, 88(6), 1430–1439.
Ertlen, D., Schwartz, D., Trautmann, M., Webster, R., Brunet, D. (2010). Discriminating
between organic matter in soil from grass and forest by near-infrared spectroscopy. Eur. J.
Soil Sci., 61(2), 207–216.
Feiss, T., Horen, H., Brasseur, B., Lenoir, J., Buridant, J., Decocq, G. (2017). Optimal
sampling design and minimal effort for soil charcoal analyses considering the soil type
and forest history. Veg. Hist. Archeobot., 26(6), 627–637.
Ferguson, D.K. (2005). Plant taphonomy: Ruminations on the past, the present, and the
future. Palaios, 20(5), 418–428.
Kögel-Knabner, I. (2002). The macromolecular organic composition of plant and microbial
residues as inputs to soil organic matter. Soil Biol. Biochem., 34(2), 139–162.
Lavelle, P., Spain, A., Blouin, M., Brown, G., Decaens, T., Grimaldi, M., McKey, D.,
Mathieu, J., Velasquez, E., Zangerlé, A. (2016). Ecosystem engineers in a self-organized
soil: A review of concepts and future research questions. Soil Sci., 181(3–4), 91–109.
Nelle, O., Robin, V., Talon, B. (2013). Pedoanthracology: Analysing soil charcoal to study
Holocene palaeoenvironments. Quatern. Int., 289, 1–4.
Peterken, G.F. (1996). Natural Woodland: Ecology and Conservation in Northern Temperate
Regions. Cambridge University Press, Cambridge.
Robin, A.-M. (2005). Épisodes majeurs de la podzolisation en forêt de Fontainebleau
(France). Essai de synthèse à l’aide du radiocarbone. C.R. Geosci., 337(6), 599–608.
Robin, V., Rickert, B.-H., Nadeau, M.-J., Nelle, O. (2012). Assessing Holocene vegetation
and fire history by a multiproxy approach: The case of stodthagen forest (northern
Germany). Holocene, 22(3), 337–346.
Robin, V., Bork, H.-R., Nadeau, M.-J., Nelle, O. (2014). Fire and forest history of central
European low mountain forest sites based on soil charcoal analysis: The case of the
eastern Harz. Holocene, 24(1), 35–47.
Sabatini, F.M., Burrascano, S., Keeton, W.S., Levers, C., Lindner, M., Pötzschner F.,
Verkerk P.J., Bauhus J., Buchwald E., Chaskovsky O. et al. (2018). Where are Europe’s
last primary forests? Divers. Distrib., 24(10), 1426–1439.
Šamonil, P., Král, K., Hort, L. (2010). The role of tree uprooting in soil formation: A critical
literature review. Geoderma, 157(3), 65–79.
Slessarev, E.W., Lin, Y., Bingham, N.L., Johnson, J.E., Dai, Y., Schimel, J.P., Chadwick,
O.A. (2016). Water balance creates a threshold in soil pH at the global scale. Nature,
540(7634), 567–569.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Soil Archives 97

Stinchcomb, G.E., Driese, S.G., Nordt, L.C., DiPietro, L.M., Messner, T.C. (2014). Early
Holocene soil cryoturbation in northeastern USA: Implications for archaeological site
formation. Quatern. Int., 342, 186–198.
Trendel, J.M., Schaeffer, P., Adam, P., Ertlen, D., Schwartz, D. (2010). Molecular
characterisation of soil surface horizons with different vegetation in the Vosges massif
(France). Org. Geochem., 41(9), 1036–1039.
Yaalon, D.H. (1971). Soil-forming processes in time and space. In Paleopedology. Origin,
Nature and Dating of Paleosols, Yaalon, D.H. (ed.). Israel University Press, Jerusalem.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
8

Continuous and Nested


Time in Historical Ecology:
Application to Soil Studies
Damien ERTLEN
University of Strasbourg, France

8.1. Interdisciplinarity and time in historical ecology

Historical ecology (HE) is about interdisciplinarity (Szabó 2015). Anthropology,


archaeology, ecology, ethnography, geography, geology and history are the
disciplines most commonly found within the framework of HE. The combination
and relative importance of these disciplines in HE vary somewhat between authors
and depend on how these disciplines are defined and delimited. Moreover, HE can
be considered as “deep interdisciplinarity” because it is not simply a temporary
collaboration or a discipline assisting another by providing a different method or
having a different point of view. Indeed, the researchers involved in HE share a
common set of questions that require addressing in a research project (Balée 2006).
Studying the dynamic relations over time between humans, “our species”, and
nature, “our environment”, is one means of formulating the central research question
(Balée 2006; Crumley 2007; Szabó 2015). This human–environment interaction is
also often considered as a central question for geographers (Dauphiné 2003). The
assumption that all environments on Earth have been affected by humans (Balée
2006; Rostain 2016) is widely accepted by HE researchers. This idea provides an
overarching dimension to any project.

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
100 Historical Ecology

Several disciplines involved in HE deal intrinsically with time: geology through


the Earth’s history, biology via the history of life and archaeology and history
involving the history of societies. Time is a common and relevant topic for
practicing interdisciplinarity. All sciences that focus on the environment eventually
became historical sciences, as we cannot understand the present without
understanding the past (Legeay 2000). This position is increasingly accepted,
particularly in ecology, because time coupled with space represents one of the
fundamental axes shaping ecological processes (Wolkovich et al. 2014). White
(2007), for example, demonstrated the necessity of considering time and various
timescales in landscape ecology and landscape management/restoration. She
explains how a lack of historical perspective or the use of an improper timescale can
lead to a misunderstanding of landscape dynamics. Crumley (2007) explains how
failing to consider time led some American ecologists to an impasse, as they
continued to look for “pristine” ecosystems lacking a human influence. In a similar
vein, Erickson (2020) describes how the “discoverers” of pristine forest in
Amazonia are very often frustrated when visiting the forest together with
archaeologists, who quickly find evidence of a past human presence. The emergence
of climate change as an issue has further challenged researchers across multiple
disciplines to better understand processes over a longer temporal scale.

The author of this chapter is a geographer … talking about time. Indeed, work on
dynamic processes (social or natural) involves more fields than the traditional
disciplines that study the past. The lack of a strong temporal framework in
geography likely makes it easier for geographers to switch from one timescale to
another, similar to how geographers traditionally act in regard to spatial scales.
Timescale issues and errors obtained when using the inappropriate timescale have
been previously identified by Rymer (1979). Time in HE has been conceptualized
by Sinclair et al. (2018) who placed a specific focus on complex systems and the
necessity to move beyond the linear, Newtonian definition of time. Here, I focus on
two practical means for approaching time with the goal of improving
communication between the practitioners of HE. To provide an example of a study
tool, I rely on the use of soils and their temporal organization in HE studies.

8.2. Continuous time

Continuous time is used by historians and the historical sciences. It is also the
vision of time learned at school. For the geologist, the Earth’s historian, it begins
with the Big Bang (ca. 4.5 billion years BP). Time is then divided by a hierarchical
and universal system into aeons, eras, periods, epochs and stages (Figure 8.1). The
limits between these time entities are often linked to life on Earth; for example, the
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Continuous and Nested Time in Historical Ecology 101

start of the second aeon, the Archean, is associated with the first record of life on
Earth. The Cretaceous is associated with the appearance of flowering plants. The
Quaternary, until 2009, was a period associated with the appearance of Hominidae;
however, following the discovery of older hominids, the Quaternary was
downgraded to an epoch, forming part of the Cenozoic era. Other transitions
between geological subdivisions are linked to changes in atmospheric chemistry and
geological events.

Figure 8.1. History of human–environment relations: a continuous-time perspective

Similarly, historians divide time into a series of successive periods. This system
is not as universal as that for geology; however, traditional periods, such as
prehistory, the Neolithic Age, antiquity and the Middle Ages, are widespread and
globally well understood, despite regional differences in terms of the absolute limits
of these periods. This subdivision of time serves not only as a tool to facilitate the
organization of Earth or human history, but often it also serves to organize
university departments and research projects. Consequently, specialists of one
period (one slice in time) develop their own community, which can sometimes be
disconnected from other communities. Each community develops its own methods,
and these closely knit research spheres are often disconnected from other periods of
time – even including the present. This means of dividing time also adds to the
difficulty for interdisciplinary communication, given that the timescales of interest
for each discipline do not necessarily match and are often difficult to bridge
(Figure 8.1). Crumley (2007) underlines the challenge in bringing together these
different systems within an “integrated framework” dedicated to HE. On the other
hand, these chronology-based divisions cannot be neglected because they remain the
most common temporal divisions used by HE researchers.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
102 Historical Ecology

8.3. Nested time

An alternative way to divide time is to consider nested portions of time rather


than continuous time slices. We can use the image of Russian dolls, where each doll
represents a different timescale. The above-mentioned geological divisions can also
be viewed as a nested timescale system because it is a hierarchical structure. In the
approach described in this chapter, however, all timescales end (or begin) at the
same point, which is the present (“Today” in Figure 8.2). This point is important
because although some archives such as sediments may have stopped accumulating
information at an earlier moment in time, we nonetheless study these archives in the
present and must consider the amount of time that the archive has been preserved
and the degradation of this archive over time. This view agrees with one of the
primary objectives of HE, which is to explain present-day landscapes and
ecosystems from the viewpoint of these multiscale temporal and spatial dynamics.
The French historian Braudel (1980), developing short- and long-term history,
provided a precursor to this nested time approach; however, it was limited to a
centennial timescale.

Figure 8.2. History of human–environment relations: a nested-time perspective


Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Continuous and Nested Time in Historical Ecology 103

We could view very short timescales, at an hourly or daily scale, as the smallest
scales of interest (smallest doll); in practice, however, these very short-term scales
are rarely used in HE. Data at a yearly or decadal scale are much more common in
HE and are described as short-term history. (In practice, use of these short-term and
long-term denominations alone is quite challenging because their magnitude can
vary between disciplines and researchers.)

Centuries and millennia are also appropriate timescales in studies of ecological


history. We can also ask what is the largest doll and how far back does HE reach
into the past. Szabó (2015) distinguishes three forms of past: the past sensu lato, the
human past and that part of the human past for which written records are available.
A consensus is emerging that HE is the history of the human–environment
relationship; therefore, it does not become necessary to extend beyond this time
frame. Some older elements, however, particularly from geology or biology, are
often useful for HE, not to study their own particular histories per se, but because
they are inherited and are implicated in the human–environmental interactions.
Crumley (2007) defends the idea that temporal ranges are limited only by what data
are available and thus relevant timescales for a study can range from very recent to
geological timescales. The development of agriculture has also been viewed as a key
point in human history and sometimes as a starting point for HE. Consequently,
post-Neolithic, or more generally the Holocene, has been considered as the time
frame for HE. However, the controlled use of fire and other forms of plant
management have been described as occurring long before crop cultivation and
neolithization.

The framework of HE provides an intellectual “contact zone” between


disciplines (Pratt 1991 in Meyer and Crumley 2012). To be effective, the contact
must concern an object of study occurring at the same time or with overlapping
timescales. With nested time, the different timescales always have a minimum area
of overlap. Describing this nested system is relatively simple using an ordering of
chronological scales, although in practice, HE researchers have various strategies
and constraints when selecting the appropriate scales for study. In the following
sections, we describe these practices and emphasize the interest in combining
different timescales to strengthen the conclusions of HE studies.

8.4. Different disciplines, different tools

Even if there were to be an eventual agreement among researchers within the


interdisciplinary consortium of HE on this necessity of using a nested time
approach, researchers would still have to agree on the magnitude and limits of these
temporal scales. Clearly, the temporal magnitude varies among disciplines. In some
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
104 Historical Ecology

cases, as described for continuous time, this is often a question of habit or discipline-
specific methods; it is not difficult to convince colleagues to switch timescales when
we demonstrate the advantages of the novel method, the relevance of the new scales
and the attraction of newly available data. In most cases, however, it is not simply a
question of methodology, but rather this method being determined by the tool.
Nearly all the tools are restricted in their magnitude of time, and below, we provide
three examples to highlight this point.

(i) Old maps and old aerial photographs are highly useful for reconstructing past
environments and landscapes. Geographers and historians are quite familiar with the
use of such tools. Increasingly, map and photograph collections are being scanned,
digitized and made widely accessible within databases. They often provide clear and
precise information in relation to the spatial distribution of forests, grasslands,
agricultural plots, etc. These tools commonly provide detailed information about
vegetation type. In Europe, maps from the 18th century begin to provide sufficiently
precise and reliable information. From the mid-19th century, maps become detailed,
clear and easy to digitize. By the 20th century, maps – coupled with the arrival of
aerial photographs – are increasingly available, and the more frequent surveys of a
given land parcel make it possible to reconstruct landscape dynamics at a higher
spatio-temporal resolution. Maps produced prior to the 18th century are available;
however, they are difficult to align, the information contained within is often
unreliable and the mapped areas are dispersed in space and time. Consequently, the
interest in using old maps markedly falls when changing from the timescale of the
last three centuries to a millennial timescale. In this latter case, a historian or a
paleoecologist must rely on other better-adapted tools to access older archives at this
longer timescale.

(ii) The Neolithic is generally now rejected as a starting point for HE.
Nevertheless, the social changes and, in particular, the arrival of a novel sedentary
lifestyle produced a massive change in the quantity and readability of archaeological
remains. Consequently, there is a wide methodological gap between pre- and
post-Neolithic sites in terms of their excavation and study. This gap also involves
temporal organization and time resolution; for example, the relative chronology built
from Neolithic pottery can approach a time resolution of a single human generation
(30 years), a very high resolution for the mid-Holocene. The chronologies of earlier
periods are based on a few pieces of stone tools and scattered radiocarbon dates. In
contrast, the material available in the Neolithic, both for relative and absolute
chronologies, is much more abundant. Therefore, the dating of sediments by
paleoecologists is greatly more critical prior to the Neolithic. A collaboration
between a paleoecologist and an archaeologist can offer completely different
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Continuous and Nested Time in Historical Ecology 105

experiences depending on the archaeologist’s temporal speciality, for example,


Neolithic versus Paleolithic.

(iii) The soil archives presented by Brasseur et al. (see Chapter 13) and also
discussed below have time limits that vary according to their location. In temperate
Europe, for example, it is accepted that all soils are younger than the last
deglaciation. During the last glacial period, which attained a maximum extent
20,000 years BP, soils were destroyed either directly by ice sheets and glaciers over
northern Europe or by periglacial erosional processes occurring in the periphery of
the ice. Consequently, a European soil’s environmental memory is restricted to the
post-glacial period. Soil scientists and paleoecologists can jump to an earlier
timescale by focusing on aeolian sediments or relict paleosoils, although this
requires a very different approach.

8.5. Examples of nested and continuous time: soils and strata

Soil lies at the earth–atmosphere interface and is studied by soil scientists


(pedologists). Soils are commonly used in HE because they store information about
past environments (i.e. Koerner et al. 1997). They are complex and open systems
under the influence of geochemical and biological processes, and soils are often
misunderstood by researchers from other disciplines (Schwartz 2012). One major
source of confusion occurs between the notions of strata, which are the elementary
bodies of sedimentological stratigraphy, and horizons, which are the elemental
bodies of soil. A major difference between these two types of layers, which
sometimes have a similar appearance to inexperienced observers, is the organization
of time.

The organization of strata is well known and follows the model of continuous
time. Usually, time is organized vertically from oldest at the bottom to most recent
at the top, and therefore we can talk about superimposed time. Geologists and
archaeologists are very familiar with this concept; however, the different visions of
time between the disciplines and the different applied timescales are sometimes
problematic. It requires great effort to clearly understand timescales used by other
disciplines. The main issue arises when comparing absolute time and relative time.
On the one hand, the expanded development of radiocarbon dating and other
absolute-age dating methods, such as optically stimulated luminescence (OSL) and
in situ cosmogenic nuclides (i.e. 10Be, 26Al), has produced a greater abundance of
absolute dates from natural sediments and has improved the accuracy of the absolute
scale. On the other hand, the archaeological findings and typologies are also very
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
106 Historical Ecology

accurate and highly relevant when they are compiled into large, spatialized
databases. Matching up absolute time and relative time remains the main challenge.

Soils and soil horizons follow very different time organizations than strata. They
are the result of the redistribution and reorganization of elements through biological
and physical processes. If the soil remains unburied, these processes continue. These
still-functioning soils are appealing for HE studies because the soil cover is
(i) nearly continuous across the surface and (ii) most information contained within
the soil is local, thereby allowing the comparison of, at a specific spatial scale,
vegetation, archaeological features and other objects of interest for HE. The
temporal organization in these living soils corresponds to the nested organization
of time.

Depth in soil Chemical compound Magnitude/timescale


Top of soil Leaves and other vegetation debris Years
Total organic matter Decades
Upper
20 cm/surficial Lignin, bacterial hexosamines, glucosamines* Decades
horizon
n-alkanes, saccharides* Century
Total organic matter Centuries to millennia
Deep horizons n-alkanes Millennia
Glycerol dialkyl glycerol tetraethers (GDGTs) Millennia
*According to Schmidt et al. (2008).

Table 8.1. Organic matter components and their respective transit times

Some elements accumulate in soils; however, most elements are in transit in


soils, and this transit time differs greatly among elements. The most resistant
mineral grains, such as quartz, can remain in soils for several million years before
being weathered. In contrast, some organic components remain in the soil for only a
single season or a few days. This transit time varies greatly, even by an order of
magnitude, among the various types of organic molecules (Table 8.1). Radiocarbon
dating of soil organic matter, an approach commonly used by HE researchers, has
been important in soil studies since the 1980s (Becker-Heidmann and Scharpenseel,
1992). Because this organic matter is heterochronous, it is exceedingly difficult to
determine its absolute age. Rather, soil scientists determine a mean residence time
(MRT, Figure 8.3) for this organic matter. This MRT depends on depth, soil type
and multiple interdependent natural variables, including temperature, pH, microbial
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Continuous and Nested Time in Historical Ecology 107

activity and vegetation type. MRT is also influenced by soil management strategies
and human-related soil disturbance, thereby making MRT extremely relevant in HE.

Figure 8.3. Total soil organic matter mean residence time (MRT) from a large set of
soils that includes various soil types under different environmental conditions

8.6. Conclusion

Historical ecology, as a science using the past to understand the present, cannot
avoid questioning time and timescales. Disciplines differ in their approaches for
dividing time. Here, we propose synthesizing these various means into two main
approaches: continuous and nested time. Although both approaches have their
unique advantages and limitations, nested time is best adapted for building a strong
and comprehensive dialogue between the disciplines working in HE. Nevertheless,
this approach does not eliminate the interdisciplinary differences and most
researchers remain narrow specialists within a single discipline. For constructive
dialogue within the framework of HE, however, we recommend that researchers pay
particular attention to the chronological views of closely related sciences to develop
their own knowledge of these alternative visions of time.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
108 Historical Ecology

8.7. References

Balée, W. (2006). The research program of historical ecology. Annu. Rev. Anthropol., 35,
75–98.
Becker-Heidmann, P. and Scharpenseel, H.W. (1992). Studies of soil organic matter
dynamics using natural carbon isotopes. Sci. Tot. Env., 117–118, 305–312.
Braudel, F. (1980). On History. University of Chicago Press, Chicago.
Crumley, C.L. (2007). Historical ecology: Integrated thinking at multiple temporal and spatial
scales. In Issues and Concepts in Historical Ecology: The Past and Future of Landscapes
and Regions, Crumley, C., Lennartsson, T., Westin, A. (eds). Cambridge University
Press, Cambridge.
Dauphiné, A. (2003). Les théories de la complexité chez les géographes. Anthropos, Paris.
Erickson, C.L. (2020). Foreword. In Methods in Historical Ecology: Insights from Amazonia,
Odonne, G. and Molino, J.-F. (eds). Routledge, London.
Koerner, W., Dupouey, J.-L., Dambrine, E., Benoît, M. (1997). Influence of past land use on
the vegetation and soils of present day forest in the Vosges mountains. J. Ecol., 85,
351–335.
Legeay, J.-M. (2000). Les temps de l’environnement. In Les temps de l’environnement,
Barrué Pastor, M. and Bertrand, G. (eds). Presses Universitaires du Mirail, Toulouse.
Meyer, W.J. and Crumley, C.L (2012). Historical ecology: Using what works to cross the
divide. In Atlantic Europe in the First Millennium BC: Crossing the Divide, Moore, T.
and Armada, X.L. (eds). Oxford Scholarship Online, Oxford.
Rostain, S. (2016). Amazonie : un jardin sauvage ou une forêt domestiquée. Actes
Sud/Errance, Paris.
Rymer, L. (1979). Historical ecology and environmental conservation. Env. Conserv., 6,
199–200.
Schmidt, M.W.I., Torn, M.S., Abiven, S., Dittmar, T., Guggenberger, G., Janssens, I.A.,
Kleber, M., Kögel-knabner, I., Lehmann, J., Manning, D.A. et al. (2008). Persistence of
soil organic matter as an ecosystem property. Nature, 478, 49–56.
Schwartz, D. (2012). Les temps du sol : interprétations temporelles de l’archivage
pédologique dans les approches paléoenvironnementalistes et géoarchéologiques. Etude et
gestion des sols, 19, 50–66.
Sinclair, P., Moen, J., Crumley, C.L. (2018) Historical ecology and the longue durée. In
Issues and Concepts in Historical Ecology, Crumley C.L., Lennartsson, T., Westin, A.
(eds). Cambridge University Press, Cambridge.
Szabó, P. (2015). Historical ecology: Past, present and future. Biol. Rev., 90(4), 997–1014.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Continuous and Nested Time in Historical Ecology 109

White, K.A. (2007). Invoking the ghosts of landscapes past to understand the landscape
ecology of the present… and the future. In Issues and Concepts in Historical Ecology:
The Past and Future of Landscapes and Regions, Crumley, C., Lennartsson, T., Westin,
A., (eds). Cambridge University Press, Cambridge.
Wolkovich, E.M., Cook, B.I., Davies, T.J. (2014). Progress towards an interdisciplinary
science of plant phenology: Building predictions across space, time and species diversity.
New Phytol., 201(4), 1156–1162.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
9

The Analysis of Relic Charcoal


Kilns for the Assessment of
Forest Trajectories
Vincent ROBIN1, Alexa DUFRAISSE2 and Claudia OLIVEIRA1
1
University of Lorraine, Metz, France
2
Muséum National d’Histoire Naturelle, Paris, France

9.1. Introduction

The demand for a reliable supply of energy to meet industrial developments has
continuously increased from the Bronze Age onwards. Energy was supplied
primarily from charcoal produced after woodland biomass exploitation until the rise
of fossil energy use during the 19th century (Pain 2017). Such biomass exploitation
for charcoal production represented an anthropogenic disturbance to the dynamics of
the forest ecosystems. Therefore, historical charcoal production has had a significant
influence on past forest trajectories, from punctual and local, to large and long
lasting, with, moreover, possible heritages on the present-day state of forests
(Ludemann 2010; Bonhage et al. 2020; Máliš et al. 2020). Thus, the reconstruction
of historical charcoal production regimes (e.g. magnitude, intensity, frequency, etc.)
might provide key insights to improve our understanding and knowledge about the
historical ecology of forest ecosystems and how it has been used as resource
corollary of human development.

Valuable data can be obtained from specific archives of past charcoal production
that are the remaining charcoal production structures (i.e. kilns) and charcoal pieces

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
112 Historical Ecology

remains (Nelle 2003; Gocel-Chalté et al. 2020). The carbonization of woodpiles


(Figure 9.1a) in soil pits during antiquity and in mound-shaped kilns from the
Middle Ages onwards (Figure 9.1b) implies that there has been an on-site
accumulation of ash and charcoal pieces. These remains, mixed with on-site burnt
soil, constitute a charcoal-rich layer (Figure 9.1c), which represents an archive that
can be used to investigate historical forest ecology. However, pit kilns are hardly
identifiable in modern landscapes without specific excavation (Deforce et al. 2021).
Therefore, investigations of charcoal-rich layers are conducted primarily on
mound-shaped kilns. These are commonly pointed out across the globe, largely
because they are fairly easy to identify from the relict platform on which the
mound-shaped kilns were originally set up (Figure 9.1d).

Figure 9.1. Various states of mound-shaped charcoal kilns: setting up the woodpile
mound (a); during carbonization (b); relict charcoal-rich layer view along a
longitudinal soil profile (c) and ongoing charcoal sampling of a relict charcoal kiln
platform (d). For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip

9.2. Looking at the platform of the kiln

9.2.1. Looking at the dimensions of the kiln platforms

An analysis of the dimensions of charcoal-rich layers can provide information


about past charcoal production.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Analysis of Relic Charcoal Kilns for the Assessment of Forest Trajectories 113

First, the thickness of the charcoal-rich layer can be assessed. We can


hypothesize that the thickness of the charcoal-rich layer is related to possible
variations in the quantity or duration of charcoal production. However, to date, no
relationship has been identified between the production processes and the thickness
of the charcoal-rich layer. Even when high-resolution indicators are used to
investigate the charcoal-rich layer, such as micro-morphological features and
paleo-temperature records (Dupin 2018), the charcoal layer appears to be a
homogeneous layer of varying thickness. The micro-morphological analyses of
some charcoal-rich layers have revealed that the charcoal makers reworked the
surface of the platforms between carbonization of the woodpiles (Gebhardt 2007;
Dupin et al. 2017). This probably caused mixing of the accumulated materials with
in situ soil, which homogenized the charcoal-rich layer. This is also why the practice
of reusing kiln platforms, potentially over a long period with many cycles of
production, is barely identifiable and quantifiable. This is a crucial issue when
attempting to quantitatively investigate the past exploitation of forest biomass for
charcoal production.

Second, the dimensions of the charcoal-rich layers can be assessed by the


characterization of their surface and shape. This is achieved by on-site
measurements of the minimum and maximum diameters of relict platforms. Such
dimensional analysis has shown that the platforms created during the 19th and 20th
centuries tend to be more ovoid in shape and larger (up to ca. 200 m²) than the older
ones, which are more circular in shape and smaller (ca. 40–80 m²; Carrari et al.
2017; Dupin 2018; Gocel-Chalté et al. 2020). However, the surface and shape of the
platforms at the regional scale vary widely and it has not been possible to identify a
significant relationship between the surface and shape of the charcoal-rich layer and
the age or quantity of charcoal production.

9.2.2. Platform inventory

The quantification of the kiln platform and the characterization of its spatial
distribution might provide valuable information that could identify the historical
patterns for charcoal production and its possible impact on past and present forest
states (Ludemann 2010). Indeed, kiln frequency differences in space over a defined
area might reflect various intensities of biomass exploitation. The exploitation might
be related to specific ecosystem attributes, such as slope, or it may be related to the
presence of local industries (Gocel-Chalté et al. 2020; Schneider et al. 2020).
Various indicators might be used in the spatial analysis of kiln distribution, such as
the distribution structure through space, neighboring distance analysis and the
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
114 Historical Ecology

topographical wetness index. However, all of them rely on an inventory of kilns that
should be as accurate as possible.

At the forest scale, the kilns can be inventoried through field prospections
because kiln platforms are identifiable as specific landforms, especially in an area
with a marked relief (Nelle 2003; Schneider et al. 2020; Figure 9.1d). However, kiln
platforms are much more difficult to detect on plains because of the lack of relief
within a flat landform or in areas with dense vegetation cover. Numerical tools for
kiln detection might be of great importance in such cases. These numerical tools can
also be used to proceed an inventory of kilns at large spatial scales. First, kilns might
appear on aerial pictures as dark dots in freshly plowed crop fields. Moreover,
during the last 20 years, airborne LIght Detection And Ranging (i.e. LiDAR) has
emerged as a powerful tool for the detection and inventory of kiln platforms among
other anthropic or natural landforms (Ludemann 2012; Puech et al. 2012). The
LiDAR technique is a high-resolution “laser scanning process” that provides a set of
contact points with x, y, z coordinates. These clouds of contact points can be used to
derive different digital relief models that can highlight micro-topographical features,
including kiln platforms (Mayoral et al. 2017; Žutautas 2017; Figure 9.2).
Furthermore, the development of automatic detection tools based on deep learning
processes over the last decade (Schneider et al. 2015; Trier et al. 2016) has provided
powerful tools that can inventory kiln platforms efficiently at large spatial scales and
extrapolate the locally gathered kiln data (i.e. upscaling, Ludemann and Nelle 2017).

Figure 9.2. Digital elevation model derived from LiDAR treatments of the same study
area across the northern Vosges in France. It highlights the presence of kiln
platforms (little round forms distributed all over the pictures). Slope model (a), light
exposure model (b) and Topographic Position Index model (c)

Moreover, it is important to stress that using numerical tools to inventory kilns


reduces any under-estimation of kilns which may be observed when kilns are
inventoried using only field surveys. For example, in a recent study across the
northern Vosges in France (Gocel-Chalté et al. 2020), both field and LiDAR
inventories were independently created. A comparison between the two revealed
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Analysis of Relic Charcoal Kilns for the Assessment of Forest Trajectories 115

that field inventories under-estimate the kiln frequency by about 30%, while the
LIDAR inventory valid detection rate for kiln platforms was 96%.

9.3. Looking at the charcoal pieces

9.3.1. Sampling

Besides the dimensions of the kiln platform, the charcoal remains may also
provide key insights. First, charcoal pieces from the charcoal-rich layer are sampled.
For one kiln, the sampling consists of randomly collecting charcoal pieces from
various depths in the charcoal-rich layer (Figure 9.1c) on manually opening small
pits (e.g. 40/40 cm). The sampling can be done by on-site handpicking or by taking a
volume of sediment from the charcoal-rich layer and sieving it to extract the
charcoal pieces. At least 100 charcoal pieces per kiln are usually collected for the
dendro-anthracological analyses. To avoid the risk of falsely recording
over-represented taxa, the sampling must be done at different locations on the kiln
platform (Nelle 2003). Classically, samples are taken in the center of the kiln
platform and on the downstream side of the platform, which is usually richer in
charcoal than the kiln itself.

For a given area under investigation, the number of kilns that must be sampled to
achieve an appropriate level of representativeness depends on 1) the surface of the
area being investigated and 2) the frequency of kilns in the area being investigated.
In accordance with these parameters, a sampling strategy should be defined that is
based on the best compromise between the expected representativeness level of the
data and the possible analytical effort.

9.3.2. Taxonomic identification

The taxonomic identification method consists of a microscopic analysis of the


sampled charcoal assemblages (larger than 800 µm/1 mm) in each studied kiln. The
analysis relies on the fact that wood has specific microscopic anatomical structures
that are preserved after carbonization, and which correspond to specific taxa or
groups of taxa. Their identification provides the taxonomic spectrum of the ligneous
species that correspond to the sampled charcoal remains. The results of the
taxonomic identification are usually expressed as the relative values of the identified
taxa frequency in the 100 charcoal pieces sampled and identified for each
investigated kiln (Figure 9.3).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
116 Historical Ecology

Figure 9.3. Results of the taxonomical analysis of charcoal pieces assemblages from
six kiln platforms in Meuse, France (JAU 1 to 6). The results show the proportion of
identified taxa when all the analyzed charcoal fragments per kiln platform are
considered. For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip

The taxonomical identification of charcoal assemblages from kilns provides


information about the taxa that were available locally as a resource and the
composition of the forest surrounding the investigated kilns. This is because the
consumed wood for charcoal production would have been taken from the area
surrounding the kiln and was dependent on the topography and the availability of
local biomass (Knapp et al. 2013; Dupin 2018). Thus, specific species selection for
the charcoal production is rare, even if hardwood of strongly re-sprouting
tree species are usually the preferential resource for charcoal production
(Grenouillet-Paradis 2012). It has been shown that charcoal can be largely produced
from softwood when this is the only available resource (Deforce et al. 2013). It has
been also shown that the kiln charcoal assemblages at the catchment scale strongly
reflect the potential forest tree distribution, which is related to environmental
conditions (Gocel-Chalté et al. 2020). For example, the taxonomic information
about the charcoal assemblages might provide evidence of forest stand disturbance
because light-demanding pioneer tree species were recorded in the sample (e.g.
Betula, Pinus; Larsen et al. 2016). Such information about past forest dynamics
becomes even more interesting when combined with structural reconstructions of the
harvested forest stand and related past forest management regimes (e.g. Deforce and
Haneca 2015).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Analysis of Relic Charcoal Kilns for the Assessment of Forest Trajectories 117

9.3.3. Dendro-anthracology

Dendro-anthracology aims to gather the information contained in tree growth


rings. Growth rings are the concentric annual rings observable on transverse sections
of wood. Expressed as a function of time, tree-ring widths reflect growth dynamics,
which depend on intrinsic characteristics (e.g. age, tree parts, etc.), environmental
conditions (e.g. climatic events) and anthropic activity, such as forestry
practices/management (e.g. coppicing).

Dendro-anthracological measurements and data acquisition rely on the same


principles as dendrochronology, with some adaption because charcoal pieces are
incomplete, usually contain short growth rings series and come from unknown parts
of the trees (Dufraisse 2006). Thus, the application of dendro-anthracological tools
requires a minimum transverse plane size of about 4 × 4 mm and at least one whole
growth ring. Furthermore, to be reliable, the dendro-anthracological analyses must
be done on at least 50–100 charcoal pieces per kiln. The dendro-anthracology is
based on qualitative and/or quantitative observations of specific parameters, such as
heartwood/sapwood discrimination, the presence of truncated tree rings and wood
shrinkage during carbonization. The charcoal–pith distances also need to be
measured, which can be recorded in different methodological ways (i.e. tree-ring
boundaries, printed target and trigonometric calculation, and image analysis;
Ludemann and Nelle 2002; Dufraisse et al. 2020). The observation of these previous
dendro-anthracological parameters in current wood stands using dendrochronology
means that an interpretative reference grid can be established (i.e. an
anthraco-typological key; Dufraisse et al. 2018) for each tree species. This grid
is necessary to classify the charcoal pieces according to their combined
dendro-anthracological characteristics. Overall, the insights gathered from
dendro-anthracology contribute to a more informed understanding of the biomass
harvest process (e.g. type of wood, season of harvest, storage). For example, in
deciduous oaks, heartwood formation takes place in 20–25-year-old trees. Therefore,
the significant recording of heartwood in charcoal pieces from kilns provides a solid
estimation of the minimum age of the exploited trees for charcoal production and it
can be linked to forest management practices, such as coppicing (Deforce and
Haneca 2015).

9.4. Looking at the ages

Chronological data recording is a key step in the analysis of historical charcoal


production and its past consequences on forest ecosystems. It is possible to use
different chronological markers to obtain charcoal production dates.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
118 Historical Ecology

The most frequently used chronological marker is the measurement of


radiocarbon ages (Walker 2005). To date a charcoal kiln, one or several charcoal
fragments must be selected according to 1) the focus of the investigation, 2) the
gathered taxonomic spectrum at the kiln scale and 3) the gathered taxonomic
spectrum at the study area scale. It is important to select a fragment of charcoal that
allows the age of the carbonization to be dated. Indeed, long-living trees, such as
oaks of beech trees, might induce a bias in age measurement due to a potential
difference of several centuries between the tree rings that were formed earlier in the
life of the tree and those that formed shortly before the tree was harvested for
charcoal production. Those latter rings are the ones targeted to date the
carbonization. To minimize such possible “old-wood effect”, charcoal pieces from
short-lived tree species, such as birch or hazel, should be preferentially selected. To
obtain a relevant radiocarbon date, fragments of charcoal with marked tree-ring
curviness or/and with clearly visible external rings and/or sapwood can also be
selected. In addition, several charcoal pieces should be used to date a kiln platform.
This will provide a more reliable age distribution for the use of the platform. At
larger spatial scales, such as the catchment or territory scale, a set of radiocarbon
dates from different kilns should be used when assessing the chronological patterns
for charcoal production. These patterns can be obtained using probabilistic models
of combined radiocarbon dates, according to their calibration ranges (e.g. KDE
model; Bronk-Ramsey 2016).

Recently, in addition to radiocarbon dates, the use of optically stimulated


luminescence dates to identify the ages of charcoal platforms has been assessed. The
technique provides interesting and promising results that enable the age range
obtained by radiocarbon dating to be shortened (Karimi Moayed et al. 2020).
Furthermore, it is possible to evaluate the chronological patterns for charcoal
production based on the dendro-anthracological measurements. In fact, some
charcoal pieces from kilns have sufficiently visible tree rings to provide dateable
growth patterns that are comparable to those used in dendrochronology (Blondel
et al. 2018). Finally, the chronological patterns for charcoal production may be
identified by an interdisciplinary study that includes the analysis of historical written
sources, such as historical maps or forest harvest registers.

Overall, such chronological information will provide important insights into the
assessment of forest disturbance due to forest biomass exploitation for charcoal
production.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Analysis of Relic Charcoal Kilns for the Assessment of Forest Trajectories 119

9.5. Conclusion

Charcoal kiln analysis provides phyto-historical insights that enable researchers


to reconstruct the history of charcoal production and its related consequences on
ecosystems, from local to large spatial scales. Thus, the charcoal analysis of kilns
can provide information about previous forest disturbances caused by biomass
exploitation for charcoal production. Indeed, the charcoal analysis of kilns can be
used to identify disturbance regime parameters, such as type of fuel, age of the
disturbance events, magnitude, frequency, return interval and intensity. Furthermore,
kiln charcoal analysis can identify the consequences of such anthropogenic
disturbance on the forest ecological trajectories due to changes in forest composition
and structure. Recent studies have focused on the effects of historical charcoal
production on the functioning of forest ecosystems. For example, it has been shown
that the presence of a charcoal-rich soil layer changes the geochemical properties of
soils, from the past up to the present day (Hardy et al. 2017; Hirsch et al. 2017;
Bonhage et al. 2020). Such long-lasting changes in soil properties, as well as
changes in forest composition and structure, are the legacy effects of historical
charcoal production on the current state of forest ecosystems. Knowledge about
these legacies is crucial to gaining insights that will help guide the future use of
ligneous resources, such as an energy transition that could lead to an increase in
wood use for biomass energy supply.

9.6. References

Blondel, F., Cabanis, M., Girardclos, O., Grenouillet-Paradis, S. (2018). Impact of


carbonization on growth rings: Dating by dendrochronology experiments on oak
charcoals collected from archaeological sites. Quaternary International, 463, 268–281.
Bonhage, A., Hirsch, F., Schneider, A., Raab, A., Raab, T., Donovan, S. (2020). Long term
anthropogenic enrichment of soil organic matter stocks in forest soils – Detecting a legacy
of historical charcoal production. Forest Ecology and Management, 459, 117814.
Bronk-Ramsey, C. (2016). Methods for summarizing radiocarbon datasets. Radiocarbon, 59,
1809–1833.
Carrari, E., Ampoorter, E., Bottalico, F., Chirici, G., Coppi, A., Travaglini, D., Verheyen, K.,
Selvi, F. (2017). The old charcoal kiln sites in central Italian forest landscapes.
Quaternary International, 458, 214–223.
Deforce, K. and Haneca, K. (2015). Tree-ring analysis of archaeological charcoal as a tool to
identify past woodland management: The case from a 14th century site from Oudenaarde
(Belgium). Quaternary International, 366, 70–80.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
120 Historical Ecology

Deforce, K., Boeren, I., Adriaenssens, S., Bastiaens, J., De Keersmaeker, L., Haneca, K., Tys,
D., Vandekerkhove, K. (2013). Selective woodland exploitation for charcoal production.
A detailed analysis of charcoal kiln remains (ca. 1300–1900 AD) from Zoersel (northern
Belgium). Journal of Archaeological Science, 40, 681–689.
Deforce, K., Groenewoudt, B., Haneca, K. (2021). 2500 years of charcoal production in the
Low Countries: The chronology and typology of charcoal kilns and their relation with
early iron production. Quaternary International, 593–594, 295–305.
Dufraisse, A. (2006). Charcoal anatomy potential, wood diameter and radial growth. In
Charcoal Analysis: New Analytical Tools and Methods for Archaeology, Dufraisse, A.
(ed.). BAR International Series 1483, Archaeopress Ltd, London.
Dufraisse, A., Coubray, S., Girardclos, O., Nocus, N., Lemoine, M., Dupouey, J.-L.,
Marguerie, D. (2018). Anthraco-typology as a key approach to past firewood exploitation
and woodland management reconstructions. Dendrological reference dataset modelling
with dendro-anthracological tools. Quaternary International, 463, 232–249.
Dufraisse, A., Bardin, J., Picornell-Gelaber, L., Coubray, S., Garcia-Martinez, M.-S.,
Lemoine, M., Moreiras, S.V. (2020). Pith location tool and wood diameter estimation:
Validity and limits tested on seven taxa to approach the length of the missing radius on
archaeological wood and charcoal pieces. Journal of Archaeological Science Reports, 29,
102166.
Dupin, A. (2018). Caractérisation du charbonnage moderne et contemporain bisontin
(Franche-Comté, France) et de son impact sur les peuplements forestiers : le cas de la
forêt de Chailluz. PhD Thesis, Université de Bourgogne Franche-Comté, Besançon.
Dupin, A., Girardclos, O., Fruchart, C., Laplaige, C., Nuninger, L., Dufraisse, A., Gauthier E.
(2017). Anthracology of charcoal kilns in the forest of Chailluz (France) as a tool to
understand Franche-Comte forestry from the mid-15th to the early 20th century AD.
Quaternary International, 458, 200–213.
Gebhardt, A. (2007). Impact of charcoal production activities on soil profiles: The
micromorphological point of view. ArcheoSciences, 31, 127–136.
Gocel-Chalté, D., Guerold, F., Knapp, H., Robin, V. (2020). Anthracological analyses of
charcoal production sites at a high spatial resolution: The role of topographical parameters
in historical tree taxa distribution in Northern Vosges Mountains (France). Vegetation
History and Archaeobotany, 29, 641–655.
Grenouillet-Paradis, S. (2012). Etudier les forêts métallurgiques : analyses
dendro-anthracologiques et approches géohistoriques. L’exemple des forêts du mont
Lozère et du Périgord-Limousin. PhD Thesis, Université de Limoges, Limoges.
Hardy, B., Leifeld, J., Knicker, H., Dufey, J.E., Deforce, K., Cornélise, J.-T. (2017).
Long-term changes of chemical properties of preindustrial charcoal particles aged in
forest and agricultural temperate soil. Organic Geochemistry, 107, 33–45.
Hirsch, F., Raab, T., Ouimet, W., Dethier, D., Schneider, A., Raab, A. (2017). Soils on
historic charcoal hearths: Terminology and chemical properties. Soil Science Society of
America Journal, 81(6), 1427–1435.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Analysis of Relic Charcoal Kilns for the Assessment of Forest Trajectories 121

Karimi Moayed, D., Vandenberghe, D.A.G., Deforce, K., Bastiaens, J., Ghyselbrecht, E.,
Debeer, A.-E., De Smedt, P., De Clercq, W., De Grave, J. (2020). Bypassing the
Suess-effect: Age determination of charcoal kiln remains using OSL dating. Journal of
Archaeological Science, 120, 105176.
Knapp, H., Robin, V., Kirleis, W., Nelle, O. (2013). Woodland history in the upper Harz
mountains revealed by kiln site, soil sediment and peat charcoal analyses. Quaternary
International, 289, 88–100.
Larsen, A., Robin, V., Heckmann, T., Fülling, A., Larsen, J.R., Bork, H-R. (2016). The
influence of historic land-use changes on hillslope erosion and sediment redistribution.
The Holocene, 26, 1248–1261.
Ludemann, T. (2010). Past fuel wood exploitation and natural forest vegetation in the
Black Forest, the Vosges and neighbouring regions in western central Europe.
Palaeogeography, Palaeoclimatology, Palaeoecology, 291, 154–165.
Ludemann, T. (2012). Airborne laser scanning of historical wood charcoal production
sites – A new tool of kiln site anthracology at the landscape level. Saguntum: Papeles del
Laboratorio de Arqueología de Valencia, 13, 247–252.
Ludemann, T. and Nelle, O. (2002). Die Wälder am Schauinsland und ihre Nutzung durch
Bergbau und Köhlerei. Forstliche Versuchs-und Forschungsanstalt Baden-Württemberg,
Abteilung Botanik und Standortskunde, Freiburg.
Ludemann, T. and Nelle, O. (2017). Anthracology: Local to global significance of charcoal
science. Quaternary International, 457, 1–7.
Máliš, F., Bobek, P., Hédl, R., Chudomelová, M., Petřík, P., Ujházy, K., Ujházyová, M.,
Kopecký, M. (2020). Historical charcoal burning and coppicing suppressed beech and
increased forest vegetation heterogeneity. Journal of Vegetation Science, 32, e12923.
Mayoral, A., Toumazet, J.-P., Simon, F.-X., Vautier, F., Peiry, J.-L. (2017). The highest
gradient model: A new method for analytical assessment of the efficiency of
LiDAR-derived visualization techniques for landform detection and mapping. Remote
Sensing, 9(2), 120.
Nelle, O. (2003). Woodland history of the last 500 years revealed by anthracological studies
of charcoal kiln sites in the Bavarian Forest, Germany. Phytocoenologia, 33(4), 667–682.
Pain, S. (2017). Power through the ages. Nature, 551, 134–137.
Puech, C., Durrieu, S., Bailly, J.-S. (2012). Aireborne LIDAR for natural environments:
Research and applications in France. Revue française de photogrammétrie et de
télédétection, 200, 54–68.
Schneider, A., Takla, M., Nicolay, A., Raab, A. (2015). A template-matching approach
combining morphometric variables for automated mapping of charcoal kiln sites.
Archaeological Prospection, 22, 45–62.
Schneider, A., Bonhage, A., Raab, A., Hirsch, F., Raab, T. (2020). Large-scale mapping of
anthropogenic relief features – Legacies of past forest use in two historical charcoal
production areas in Germany. Geoarchaeology, 35(4), 545–561.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
122 Historical Ecology

Trier, Ø.D., Salberg, A.-B., Holger Pilø, L. (2016). Semi-automatic mapping of charcoal kilns
from airborne laser scanning data using deep learning. In Oceans of Data, Matsumoto, M.
and Uleberg, E. (eds). Archeopress Publishing LTD, Oxford.
Walker, M. (2005). Quaternary Dating Methods. John Wiley & Sons, Chichester.
Žutautas, V. (2017). Charcoal kiln detection from LiDAR-derived digital elevation models
combining morphometric classification and image processing techniques. Master thesis,
University of Gävle, Gävle.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
10

Ancient Trees and Botanical


Indicators as Evidence for
Change and Continuity in
Landscape Evolution
Ian D. ROTHERHAM
Sheffield Hallam University, UK

10.1. Introduction

There is considerable interest in the idea of either “ancient” woods or “old-growth”


forest (e.g. Rackham 1976, 1980, 1986, 2007; Peterken 1981, 1996) and in relating
this to contemporary management. The conceptual basis of this approach to treed
landscapes emerged in the United Kingdom in the 1970s through discussions between
Cambridge University scholar Oliver Rackham and government agency woodland
advisor George Peterken (Peterken pers. comm. and 2018). These early discussions
were a first attempt to rationalize the place of woodlands and “woods” in the historic
landscape. Additionally, evidence emerged that these landscapes with long timelines
were “ancient” (defined by Rackham as pre-1700 AD or Peterken as pre-1600 AD),
but not “primeval” wildwood. These were anciently treed countryside with long
histories of human usage and management but ancient woods were not de facto
“wildwood” (Rotherham 2007). With wider awareness of the importance of ancient
woodlands, the misconception that these equated to wildwood grew. Indeed, many
approaches to wooded landscapes and their management were fundamentally flawed
by lack of understanding of the historic context of sites and their ecologies (Rotherham
2011). Research in the United Kingdom and across Europe addressed these issues to

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
124 Historical Ecology

provide a robust interrogation of forest and woodland dynamics and better inform
contemporary conservation management (Çolak et al. 2018). This chapter presents a
brief overview of evidence for these assertions and summarizes a conceptual model
into which historical and ecological information can be placed. This allows critical
assessment of woodland antiquity and ecological continuity.

10.2. What is ancient woodland? Questions of woods versus


old-growth forest, and of continuity versus antiquity

Taking “ancient woodland” in England as defined in England, for example, it is an


area wooded continuously since at least 1600 AD, though some definitions give 1700
AD. The choice of this date is explained in Peterken (1977, 2018). Definitions are of
course essential but inherently problematic. “Continuously wooded” in this sense does
not imply continuous physical cover of trees and shrubs since open land (temporary and
permanent) is an important part of a woodland landscape. Furthermore, in all ancient
woods, trees and shrubs have been periodically felled and if an area remained woodland,
through either replanting or regrowth of, for example, coppiced areas, it is still ancient.

Indeed, and potentially confusing (see Rotherham 2007) if a working wood has
been cut-over many times the ancient woodland may have few old standard trees,
but may have very old (frequently unrecognized) coppice stools. Guidance in the
United Kingdom suggests that if over the last four centuries, woodland has
experienced a long period when the land was effectively open (e.g. grassland, heath,
moor or arable), then it is “recent” woodland. Such a site may have high nature
conservation value, but is not ancient woodland. However, reinterpretation of
site-based studies and criteria from an ecological historical viewpoint has provided a
more nuanced interpretation of these landscapes (Rotherham 2017) that has
implications for the Vera (2000) vision of the European landscape. Multidisciplinary
studies have shown how ancient woods have their origins in the enclosure of land from
the wider countryside of wood-pastures during the early medieval period. In England,
this was associated with the imposition of Norman feudal approaches following the
Norman Conquest (1066) and documented in the Domesday account (1086).
Additionally, it is apparent that significant areas of unenclosed wood-pastures remain
unrecognized in the modern countryside and, through abandonment or degradation,
have been largely wrongly assigned by contemporary ecologists.

10.3. The value of ancient woods

Ancient woodlands are of very high ecological and landscape importance and are
also hugely significant for their archaeology, historical significance and heritage
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Ancient Trees and Botanical Indicators as Evidence for Change 125

(see Rotherham et al. 2008). Numerous rare and often threatened species are
associated with ancient woods (Peterken 2000), and relatively undisturbed sites may
contain important features of historical, archaeological and landscape significance.

Woodlands and their trees form one of the most valuable habitats for wildlife
across Britain and Europe (Box 10.1). In lowland situations, an ideal scenario for
nature conservation would be mature, mixed woodland including ancient and
veteran trees, dead branches and timber lying on the ground, open rides, clearings
and glades scattered throughout the wood, with diverse ground vegetation and
recognizable field layers of shrubs with no single component dominating. This
combination of habitats within the woodland macro-habitat offers the greatest range
of conditions to attract the widest variety of wildlife.

1) Exceptionally rich in wildlife with rare species and habitats.


2) Provide high-quality renewable resources of hardwood timber and other woodland
products.
3) May hold surviving descendants and features from natural “forests”.
4) Reservoirs from which wildlife can spread into newly created woodlands.
5) Integral components of historic landscapes.
6) May have historic features little altered by modern cultivation or macro-disturbance.
7) Contribute to sense of place and imagination for a locale.
8) Importance as features in the landscape.

Box 10.1. The conservation value of woods and anciently treed landscapes

10.4. Methodology

This chapter considers (i) methodological issues – approaches to survey,


information gathering and analysis (field survey, citizen science, GIS manipulation,
LiDAR analysis), (ii) review of literature (indicators, colonization rates, tree forms
and aging), (iii) emerging concepts (shadow woods and indicators) and (iv)
synthesis of on-going research from case study examples.

10.4.1. Evidencing ancient woodlands and the use of indicators

Identifying ancient woodlands and confirming their status is important with


countryside under pressures from urban and agricultural or forestry development.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
126 Historical Ecology

Established practice to identify ancient woodland mixes map-based and


field-evidence such as indicators. The latter are species, usually vascular plants,
more commonly found in ancient woods than in recent sites. This is evidenced from
regional site/species lists. The occurrence of suites of indicator species is used as
evidence for woods being ancient, with interpretation interrogated at local and
regional levels. No individual plant species is a perfect ancient woodland indicator,
and in particular, the degree of association of species with ancient woodland varies
across the countryside with geology, topography, climate and land-use factors.

The variety of plants and trees within woodlands depends on soil type and
condition and also on sunlight reaching the woodland floor. So, for example, heavily
shaded woodland (such as coniferous monocultures and beech-dominated woods) is
not conducive to a diverse ground flora. Without diverse flowering plants,
invertebrate variety is reduced and populations of small mammals and of birds may
be restricted. Beech trees come into leaf before oaks and also cast a dense shade so
there is hardly any ground layer under their canopy. Generally, mixed composition
and age of woodland are necessary for high nature conservation value (Peterken
1981). This structure includes clearings and glades formed by tree fall or loss of
major limbs from large mature trees. When applied systematically over a long
period, traditional woodland management such as coppicing creates varied light
conditions. Recently coppiced compartments are opened to sunlight, whereas later in
the cycle, under re-grown springwood conditions are much darker. The impact of
management history has major implications for the likelihood of indicator plant
presence and visibility. It also affects invertebrates and other indicators both
indirectly through vegetation and directly through microclimatic conditions.

Coppicing involves cutting and removing “wood” from the woodland


environment. In the past, this included “tidying up” coppice stools and in the
working wood, having a “clean” housekeeping approach with clearance of dead
fallen limbs and “ramel” left after extraction. Saproxylic invertebrates (associated
with dead wood) are adversely affected by such disturbance and removal of habitat.
The stage of dead wood decay is important with some invertebrates associated with
recent dead wood and others with well-rotted material in dead standing trees.
Ground-layer flowering plants are important nectar and pollen sources as are
flowering trees and shrubs. Mature hawthorn and wild fruit trees such as crab apple
and wild pear, for instance, are important food sources for adult deadwood beetles,
craneflies and hoverflies whose larvae feed on ancient trees. Management for
traditional and industrial products such as charcoal and “white-coal” (kiln-dried
wood used for metal-smelting) also involved the stripping of vegetation and soil
from the woodland floor to cover charcoal clamps and white-coal kilns. This
resulted in intensively managed early industrial woods with very limited diversity of
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Ancient Trees and Botanical Indicators as Evidence for Change 127

flowering plants, and soils with depauperate upper horizons and sometimes replaced
entirely by a shallow layer of pure charcoal dust. Botanical indicators can inform
judgments of both woodland continuity (through presence) and past management
(through absence).

Anemone nemorosa (wood anemone)


Paris quadrifolia (herb Paris)
Lamium galeobdolon (yellow archangel)
Mercurialis perennis (dog’s mercury)
Circaea lutetiana (enchanter’s nightshade)
Conopodium majus (pignut)
Veronica montana (wood speedwell)
Lysmachia nemorum (yellow pimpernel)
Allium ursinum (wild garlic)
Luzula pilosa (hairy woodrush)
Luzula sylvatica (greater woodrush)
Oxalis acetosella (wood sorrel)
Hyacinthoides non-scriptus (bluebell)
Narcissus pseudonarcissus (wild daffodil)
Melampyrum pratense (common cow-wheat)
Corydalis claviculata (climbing corydalis)
Chrysosplenium oppositifolium (opposite-leaved golden-saxifrage)
Lonicera periclymenum (honeysuckle)
Millium effusum (wood millet)
Melica uniflora (wood melick)
Melica nutans (Mountain melick
Valeriana officinalis (common valerian)
Orchis mascula (early purple orchid)
Primula vulgaris (primrose)
Carex remota (remote sedge)
Carex sylvatica (wood sedge)
Carex pendula (pendulous sedge
Galium odoratum (woodruff)
Holcus mollis (Creeping soft-grass)
Deschampsia flexuosa (wavy hair-grass)
Festuca gigantea (giant fescue)
Geum urbanum (wood avens)
Ilex aquifolium (holly)
Corylus avellana (hazel)
Prunus padus (bird cherry)

Table 10.1. Exemplar botanical indicators of ancient woodland sites – based


on the English Peak District and Rotherham et al. (2008), Rotherham (2011,
2017). The use of indicator lists must be regionally validated
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
128 Historical Ecology

It is important to recognize the diversity of woodland types and their origins that
are broadly: medieval and industrial coppice, park and pasture woods, ancient forest,
wooded common, linear remnant and fragments (Rotherham 2007, 2017). Evidence
ancient status, and particularly the use of botanical indicators, generally relates to
former coppice woods (Rotherham 2011, 2013). Ongoing studies over 30 years,
long-term action research with expert stakeholders and detailed regional audits
allowed a review of evidence. The two-year project on the Woodland Heritage
Manual (Rotherham et al. 2008) evolved into a series of expert stakeholder
workshops examining and reviewing woodland indicators, woodland inventories and
associated landscape issues. This research reviewed the use and interpretation of
indicator species, of inventory lists and the advocacy of holistic evidence-based
evaluation.

Ancient woods are initially identified on map bases. In England and Wales, the
main cartographic sources used to identify potential ancient woodland sites were
1:25,000 maps from the 1920s and 1930s, the first edition one-inch Ordnance
Survey maps (generally mid-1800s), aerial photographs and, if available, survey
reports too. In the current work, the use of LiDAR imagery and GIS computer
mapping helps enormously. Ground-truthed where possible, most assessments of
ancient woodland status were originally from research in the late 1970s and early
1980s, and essentially intuitive lists generated by known local experts and inspection
of First Edition Ordnance Survey Maps (dating around 1830 to 1840). If available,
earlier documentation supported assessments and most sites had ecological surveys
over the following 10 to 15 years. These varied in merit depending on who the
surveyors were and why they were done. Until the 2008 Woodland Heritage Manual
(Rotherham et al. 2008), few sites had evaluations of additional factors like soils,
woodland archaeology and heritage, or historic interest.

From the review, it was clear that the broader context of historic sources and
archaeological evidence was often overlooked and even approaches such as
interpretation of map-based place-names, for example, were not widely used.

10.4.2. Tree form and growth as evidence of antiquity and continuity

Furthermore, ancient woods which lacked large, old-looking, standard trees were
frequently overlooked. Given that former medieval coppice woods, because of their
history of management inherently lack veteran standard trees, this is a serious flaw.
Additionally, there is a lack of accepted, robust approaches to aging old coppice
stools, and reluctance by many organizations to recognize these as ancient or
veteran. Yet the presence of verifiable ancient coppice stools can provide
confirmation of woodland status going back centuries. Furthermore, such trees are
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Ancient Trees and Botanical Indicators as Evidence for Change 129

evidence of both the antiquity of a site and of former management. Other trees
which are clonal or naturally self-coppicing (such as Ilex aquifolium, Alnus
glutinosa or Tilia cordata) may provide further information on the landscape
timeline.

Figure 10.1. Ancient oak coppice stool computer model reproduced with
permission from Vrška et al. (2016). For a color version of this
figure, see www.iste.co.uk/decocq/ecology.zip

Recent work in the Czech Republic (Vrška et al. (2016) suggests oak coppice
(Quercus robur) stools up to 825 ± 145 years in age (Figure 10.1), and this has
implications for dating both coppice woods and shadow wood “medusoid” oaks in
relict treescapes (Rotherham 2017). Along with the dating of small-leaved lime
coppices at well over a thousand years in, for example, Cumbria in England (Donald
Pigott pers. comm.), and the recent recognition of ancient but small veteran
specimens of hawthorn (Crataegus monogyna) and rowan (Sorbus aucuparia) in
shadow woods (Rotherham 2017), this work provides evidence of treescape
antiquity, continuity and persistence.

10.4.3. The importance of ancient and veteran trees in woodland

The presence of ancient and veteran trees in woodland adds a highly valuable
habitat component for many uncommon and rare species, particularly invertebrates
(Box 10.2).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
130 Historical Ecology

Ancient: A tree that is very old, in the declining stages of life, and usually relatively
large in girth in relation to other trees of its species. This depends on how and where it has
grown. Ancient trees are not always particularly tall but stand out visually as being very
special. Worked trees (i.e. those formerly managed to produce particular crops and
products) can be difficult either to recognize or to date. This applies especially to former
coppice trees or to some smaller ancient trees.

Veteran: A tree usually in a mature stage of life and with important wildlife and
habitat features including: hollowing or associated decay, fungi, holes, wounds and large
dead branches. These include old trees as well as younger middle-aged trees with
premature aging characteristics. These processes with species and with levels of
environmental stress, for example, upland or lowland status.

Box 10.2. “Ancient” and “veteran” trees (Rotherham et al. 2008)

10.4.4. Soils and sediments

Evidence for woodland history can be found in soils, sediments and other
deposits, and yet most ecologists overlook these. Indeed, until relatively recently,
most paleoecologists and paleobotanists dismissed wet sediment deposits in ancient
woods as being too fragmented and disturbed to be analyzed. However, current work
is showing that timelines of pollen and insect remains can be retrieved from sites to
help re-construct vegetation and faunal histories over many centuries (Helen Shaw
pers. comm.). Furthermore, such evidence adds to work on archaeological remains
which also provide information on land-use and disturbance. The replacement of
woodland soil by layers of charcoal provides insight into intensive usage for
woodland industries (Rotherham 2007). The earthworks within a wooded landscape
can give remarkable insight into human usage from modern times back to prehistory
(Rotherham 2007). These may be evidence of long periods of non-woodland status
with, for example, Roman-British or Dark Ages cultivation features, and periods of
both “coppice wood” management with woodbanks and deer park usage with
hunting structures.

In order to identify and map features such as earthworks, the application of


LiDAR (Light Detection and Ranging), aerial photography, GPS (Global
Positioning System) and drone flyover has revolutionized what is possible. With
sites identified from images they can be located and mapped on the ground in order
to be ground-truthed. From these information sources, computer-generated maps
GIS (Geographic Information System) and terrain models can also be created. These
technological tools can be combined with ecological and historical (archive)
information to provide a robust and refined assessment of a treescape site.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Ancient Trees and Botanical Indicators as Evidence for Change 131

Furthermore, data gathered on, for example, botanical indicator species distributions
can be used to create GIS “heat-maps” and thus to reconstruct possible past
landscapes from the modern data.

10.5. An emerging woodland paradigm

What we define as “woodland” and as “ancient” is not as simple as might be


assumed. Most sites recognized through both indicators and historical sources are
medieval coppice woods enclosed from former wood-pasture or wooded common.
Lack of recognition means that many sites with different timelines and histories may
be excluded from ancient status, and the consequences for conservation and
management may be serious. There are two main issues arising from the work so far.
The first is the need for robust indicators and second, for these to be placed in a
broader framework for integration and interrogation.

However, with the exceptions of obvious pollarded trees, for example, “worked”
trees, “modified” trees and naturally contorted trees (such as by animal grazing
and/or extreme weather) have tended to be overlooked. Coppice trees such as
ancient limes, alders and willows have been mostly neglected except for the
pioneering work of Donald Pigott (pers. comm.). Work on some species such as
sweet chestnut is currently in process, but there is a tendency to overlook smaller
species and peripheral habitats such as “shadow woods” (Rotherham 2017). Upland
sites and coppices are often ignored, which is problematic. Whilst understanding and
aging these specimens may be difficult, such trees are intimately tied to human
countryside exploitation and to extreme weather and climate. Many of these
specimens are older than “standard” trees, which are more generally recognized as
significant.

10.6. A simple new conceptual framework

The approaches described can be incorporated into a holistic model to bring a


multidisciplinary evidence-based approach to ancient woodland assessment
(Rotherham 2011). The basic steps are indicated in Figure 10.2.

The enhanced understanding of woodland origins, antiquity, continuity and


change can then be placed into a simple conceptual framework in relation to the
ideas postulated by Vera (2000) of the European open savanna primeval landscape
(Figure 10.3).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
132 Historical Ecology

•Ancient trees / indicator species


Ecological •Occurrence/behaviour or status/abundance
evidence

•Earthworks and associated features - prehistoric, medieval, industrial,


Field modern; archaeology 'of' the wood & 'in' the wood
archaeology •Soils & sediments, palaeo-ecology, palaeo-botany
•Veteran & ancient worked trees - copices, stubs, pollards, shreds, medusas

•Maps, archival documents, other accounts.


Historic •Field-name and place-name evidence
sources

Figure 10.2. Evidencing ancient woodland. For a color version


of this figure, see www.iste.co.uk/decocq/ecology.zip

Medieval wood-pastures
& old-growth forests

European
primeval Unenclosed
Enclosed medieval medieval wooded
coppices treescapes
commons
[Vera vision]

Modern 'ancient' woods &


'shadow woods'

Figure 10.3. Simple conceptual framework for woodland origins, continuity and
landscape change in forests. An evidence-based model derived from Rotherham
(2011, 2017). For a color version of this figure, see www.iste.co.uk/decocq/
ecology.zip
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Ancient Trees and Botanical Indicators as Evidence for Change 133

10.7. Conclusion

Recognition of “ancient woodland” diversity has implications for future


protection, conservation and management. However, the concept evolved piecemeal
to leave sites vulnerable to damage or destruction through inappropriate
development. A more holistic evidence-based approach drawing on assessments of
botanical indicators, veteran trees, soils, archaeology and historical records is more
robust. The conceptual framework is supported through a mixed-methods approach
to information gathering and evaluation. Importantly, through the conceptual
framework and the emergence of the “shadow woods” concept, we are able to place
“woods” in their historic timeline in relation to the wood-pasture and old-growth
forest origins of European treescapes (Vera 2000). Furthermore, the evidence for
both continuity and change in these landscapes is significant in developing and
applying realistic future policies to treescape management.

10.8. References

Çolak, A.H., Kirca, S., Rotherham, I.D. (eds) (2018). Ancient Woodlands and Trees: A Guide
for Landscape Planners and Forest Managers. Turkish Academy of Sciences, Ankara and
IUFRO, Vienna.
Peterken, G.F. (1974). A method for assessing woodland flora for conservation using
indicator species. Biol. Conserv., 6, 239–245.
Peterken, G.F. (1977). Habitat conservation priorities in British and European woodlands.
Biol. Conserv., 11, 223–236.
Peterken, G.F. (1981). Woodland Conservation and Management. Chapman and Hall,
London.
Peterken, G.F. (1996). Natural Woodland: Ecology and Conservation in Northern Temperate
Regions. Cambridge University Press, Cambridge.
Peterken, G.F. (2000). Identifying ancient woodland using vascular plant indicators. Br.
Wildl., 11, 153–158.
Peterken, G.F. (2018). Ancient woodland in concept and practice. In Ancient Woodlands and
Trees: A Guide for Landscape Planners and Forest Managers, Çolak, A.H., Kirca, S.,
Rotherham, I.D. (eds). Turkish Academy of Sciences, Ankara and IUFRO, Vienna.
Rackham, O. (1976). Trees and Woodland in the British Landscape. J.M. Dent & Sons Ltd,
London.
Rackham, O. (1980). Ancient Woodland: Its History, Vegetation and Uses in England.
Edward Arnold, London.
Rackham, O. (1986). The History of the Countryside. Dent, London.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
134 Historical Ecology

Rackham, O. (2007). Woodlands. Collins New Naturalist, London.


Rotherham, I.D. (2007). The implications of perceptions and cultural knowledge loss for the
management of wooded landscapes: A UK case-study. For. Ecol. Manage., 249,
100–115.
Rotherham, I.D. (2011). A landscape history approach to the assessment of ancient
woodlands. In Woodlands: Ecology, Management and Conservation, Wallace, E.B. (ed.).
Nova Science Publishers, Inc., Hauppauge, New York.
Rotherham, I.D. (2013). Ancient Woodland: History, Industry and Crafts. Shire Publications,
Oxford.
Rotherham, I.D. (2017). Shadow Woods: A Search for Lost Landscapes. Wildtrack
Publishing, Sheffield.
Rotherham, I.D., Jones, M., Smith, L., Handley, C. (eds) (2008). The Woodland Heritage
Manual: A Guide to Investigating Wooded Landscapes. Wildtrack Publishing, Sheffield.
Vera, F.W.M. (2000). Grazing Ecology and Forest History. CABI Publishing, Oxon.
Vrška, T., Janík, D., Pálková, M., Adam, D., Trochta, J. (2016). Below- and above-ground
biomass, structure and patterns in ancient lowland coppices. iForest, 10, 23–31.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
11

Towards a Methodological
Framework for Investigating the
Hidden History of Woodland
Covers
Damien MARAGE1, Catherine FRUCHART1, Isabelle JOUFFROY-
BAPICOT1, Olivier GIRARDCLOS1, Vincent BALLAND2
1
Bourgogne-Franche-Comté University, Besançon, France
2
Bourgogne-Franche-Comté University, Dijon, France

11.1. Why talk about hidden history when studying forest vegetation?

In the wake of the craze for environmental studies in history and historical
ecology, forest has become an interdisciplinary field of studies. Results in
environmental sciences are confronted with historical and archaeological data.
Nowadays, if the ecology of present forest vegetation is well documented, historical
and archaeological studies remain rare, mostly because field surveys are hampered
by the presence of vegetation and also because studying historical data is
time-consuming: archives are plethoric and scattered among many private and
public funds. Thus, our perspective on forests is, in a way, still “hidden”. This
contribution is a short review of the major disciplines and approach that we propose
to cover for the studies of historical ecology of forests. It mentions works that have
already been carried out with such approaches in France. We finally suggest a
methodological framework that could unravel the “secrets” of European forests
vegetation history over the Holocene period.

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
136 Historical Ecology

11.2. From recent forests: a synecological point of view

Reconstructing past forest dynamics presupposes an understanding of the present


forest dynamics. The study of natural forest ecosystems is a key step to achieve this
reconstruction.

Historically, the complex unraveling of ecological interactions was developed in


the classic paper on pattern and process in forest communities by Watt (1947). One
conspicuous introduction of these concepts was the review made by Whittaker
(1953) which used the Watt pattern-and-process paradigm to redefine the “climax
concept”, and furthermore to link life trait history and biogeography.

Under unconstraint conditions, the light available under the tree canopy is
reduced and the presence of the tree considerably modifies the microclimate. These
conditions determine which tree species can survive underneath the canopy. The
establishment of a new dominant canopy represents the closure of the
death/birth/death cycle that can be thought of as the typical small-scale disturbance
of a forest. According to Van Steenis (1956), tree species should be classified into
three functional groups: (i) pioneers, (ii) post-pioneers and (iii) dryads. One
characteristic of the mortality of trees in the canopy, and the associated openings
(“gap formation”), is the size of the created gaps. Species attributes are important to
differentiate the gap-size-related regeneration success of various forest ecosystems.

The persistent seed bank is a dominant driver in community assemblages,


especially in the earliest stages of secondary succession next to vegetation extant or
seed rain, for example. The seed bank remains important long after the canopy has
closed, standing as a declining input along canopy’s development and maturation.
Seeds may be recruited from the persistent seed bank in any woodland with
recurrent short-interval disturbances as coppiced stands, with forest canopy gaps or
small gaps in the herbaceous layer. The latter increasingly contributes to viewing
seed banks as a functioning part of non-successional communities in temperate
deciduous forests (Plue et al. 2010). For ecological processes, three main categories
in forest historical ecology have to be considered: (i) tree functional groups, (ii)
forest-gap dynamics and (iii) soil seed bank dynamics.

11.3. From the walls: ancient documents and maps

Textual archives and historic maps are essential to reconstruct forest landscape
history and to understand how history legacies may influence the current forest
vegetation. In France, historians have used a regressive analysis to study beyond the
French historic “Eaux et Forêts” reformation, in 1669. Ancient archives from the
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Towards a Methodological Framework for Investigating the Hidden History 137

pre-industrial period are difficult to analyze. They require special skills in


paleography and in linguistics for their deciphering and good understanding
(Figure 11.1). The history of French forests is well documented, and we shall
examine the major studies and sources.

Figure 11.1. 1474–1475, Accounting of the châtellenie de Glenne (Saône-et-Loire,


France), Gruerie section, sale of wood, Archives départementales de la Côte d’Or
B 4879-2, fol. 28 r. For a color version of this figure, see www.iste.co.uk/decocq/
ecology.zip

“Modern” maps (18–19th century) provide substantial data covering the whole
of France. The “Etat-Major” maps, produced between 1820 and 1866 on the entire
metropolitan area, are reference maps that can be used by ecologists to identify
ancient forests, as opposed to “recent” ones (Bergès and Dupouey 2017).

The “Napoleonian cadaster” was realized during the first half of the 19th century
in order to inform on tax and property data for all the French communes. The maps
are composed of thousands of separate sheets giving an accurate planimetry of land
parceling and indicating toponyms. Its matrix and tables detail land-use partly
composed of private and public wood covers. They also provide the names of
owner, plots surfaces, function and land-use also including information on the forest
nature and regime. The archives of the “Eaux et Forêts” also provide interesting data
about public forests management plans of the 19th to 20th century (Corvol 1999).
During the French Revolution, at the end of the 18th century, documents with
accurate information were produced through the sale of confiscated properties,
previously owned by ecclesiastic institutions and noble people.

The Cassini map was drawn up during the second half of the 18th century over
the entire kingdom of the time. Its main purpose was to represent the most important
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
138 Historical Ecology

traveling axes and the various inhabited places of the country. Because of its lack of
precision, it is difficult to use it for the study of woodland cover change because
forest boundaries were generally roughly represented and the restoration of forest
cover was poorly reliable, especially for areas located on reliefs or for woodland of
little importance. Apart from this map, especially for the period from the end of the
17th century to the end of the 18th century, the seigneurial and estate woods were
generally mapped and represented forest cuttings and gave precise elements on
cantonment and on the capacity and layout of the “reserve quarter” and regulated
cuts (Rochel 2017).

Additionally, textual archives written between 1661 and 1680 in relation with the
French reformation of forests provide precise data on forests condition, forest stands
and management (Poublanc 2015). These documents can help reconstructing the
past forest management, tree conditions and stands characteristics in royal forests
and, sometimes, in religious or secular community woodlands. Seigneurial accounts
from serial and annual sources, can help to reconstruct, sometimes over long
periods, forest management methods and harvesting. They inform on the frequency
of felling and the importance of grazing and are of great help in measuring the
anthropogenic activities. In some regions, it is possible to go as far as the end of the
Middle Ages like in Burgundy, for example, where we have access to the accounts
of the Duchy of Burgundy from the 14th century (Beck 2008); the parchment rolls
of the County of Savoy provide information on forest management in county forests
since the last quarter of the 13th century.

For areas that were dependent on the kingdom, or for lay or ecclesiastical
communities after the 1669 Ordinance, the documentation could be cross-referenced
or completed with the minutes of issue of the “Eaux et Forêts” cutting permits.
These documents record auction stages and methods of cutting and emptying and
control visits or resetting that were carried out, followed by logs re-examination.

The archives for private woods are less frequent. However, public archives
departments occasionally keep bundles of titles attached to ecclesiastical or secular
lordships. These collections contain various deeds, sometimes dating from the Early
Middle Ages. They can help to characterize and evaluate how forestry practices
were used by residents. Burrows, documents issued from the 15th century, give
many details about woodlands that are unrelated to the seignorial domain, but linked
to the peasantry’s censive.

Complementary information can be obtained by other sources of documents


where social aspects of the woodland landscape are involved. Court and notarial
records that settled conflicts often reveal past socio-economic frameworks and attest
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Towards a Methodological Framework for Investigating the Hidden History 139

to methods used to cut down woodland. They give an overview of the professions
that revolve around wood like logging contracts, the market of transports, wood
purchase and contracts for the processing of wood products.

11.4. From the wood: dendrochronology

The acyclical nature of droughts gives the opportunity to use them as temporal
benchmarks and build long chronologies called masters that are useful afterwards
for dating many archaeological sites. Regional master curves are based on thousands
of oaks (Quercus robur and Q. petraea), aged 1 to 200 years, discovered both in
natural deposits and in archaeological settlements. They collectively cover the entire
Holocene period since 10,430 BP. In optimal cases, felling-dates yearly identified
are linked to constructions, repairs of houses and/or villages. It pushed further into
the past the horizons of history to the Neolithic (Petrequin et al. 1998).

Dendrochronology provides variate ecological records. Mathematical tools had


made possible the study of the tree response to climate, allowing numerous
dendroclimatic reconstructions of temperatures and rainfall variabilities, at an annual
resolution (Büntgen et al. 2011). The major result was to highlight an increasing
radial growth trend during the 20th century at a global scale and provide climatic
feedback. For example, in France, this trend was identified for seven forest species
belonging to the entire panel of forests ecosystems (Badeau et al. 1996).

Growth rings also reveal forest stands characteristics. For example, radiocarbon
dating up to 500 years tauzin oak stumps (Q. pyrenaica), historically coppiced,
contrast with sprouts which only presented a few dozens of rings (Salomón et al.
2016). Tree-ring methods have been tested to detect past disturbances, such as
canopy disturbance. A study by Altman et al. (2013) established a link between
growth characteristics and past forest management, as ring-width profiles responded
to recurrence and intensity of thinning.

Dendroecological interpretation of archaeological woods provides records over a


larger timescale than with living trees. According to dendrochronological data and
historical resources, long-distance transport of valuable timber from production
areas to deforested areas was already highly developed during the Middle Ages from
countries around the Baltic Sea across western Europe (e.g. Haneca et al. 2005).
France remained at the margin of international transport, as timber was exported
from mountains to foothills. At least, timbers were floated on the river Moselle,
from the upper river Meuse and Yonne basin or from the river Durance to the Rhone
corridor.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
140 Historical Ecology

Finally, dendrochronologists describe the growth of trees exploited in the past in


order to characterize the resources needed for constructions. The shape of a tree
growth curve could provide valuable information on the levels of competition to
which it had been subjected during the different phases of its life, and therefore on
the management (or non-management) of the stand where it grew up (e.g. Billamboz
2014). For example, it is established, for several antique and medieval sites
(Girardclos and Perrault 2019), that the growth of oaks from the age of 20–30 years
was below the growth of current oaks managed in high forest and could be explained
by low forests thinning in the past and an unclear distinction between thinning and
logging phases.

The case of charcoal remains is interesting within interdisciplinary approaches.


Largest pieces give the opportunity to analyze the size of cutting and growth rings.
The remains of charcoal from industrial sites, such as charcoal kilns platforms,
which can be studied by archaeologists through spatial analysis, tend to indicate that
gathering linked to the required volume of wood is very local (Ludemann et al.
2004). The charcoal produced is eventually exported.

In these studies, the ecological information extracted from ancient woods was
obtained by comparison with the growth of living trees in known contexts (Dufraisse
et al. 2018). In principle, the hypothesis that the biological mechanisms of growth
are unchanged between ancient periods and today could be tested. Without looking
for current relics of forests from the past, the goal was to compare measurements
made on ancient trees with biological models built on living trees. There are
benchmarks that make it possible to quantify the density of stands by knowing both
the diameter and height growths.

11.5. From the ground: palynology

Pollen analysis has been used for more than a hundred years for the
reconstruction of vegetation and the study of dynamics and distribution changes in
Quaternary vegetation (Seppä 2007). Many continental sequences allowed to follow
forest change from the Late glacial (ca. 17,000 to 11,500 BP), and even more, to the
Holocene period (ca. 11,500 BP to present) with tree migrations from refugees in the
Mediterranean and the Balkans, in response to warming. Forest recolonization in
Europe followed a general trend (Figure 11.2): (i) a pioneer dynamic with pine and
hazelnut in the Early Holocene (11,500–8,200 BP), (ii) the establishment of
mesothermophilous oak forests during the Climate Optimum of the Mid-Holocene
(8,200–4,200 BP) and (iii) the development of oak-beech forest, often accompanied
by fir and/or spruce in mountain areas, at the favor of wetter and cooler climate
conditions during the Late Holocene (4,200 BP to present). Hornbeam was the last
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Towards a Methodological Framework for Investigating the Hidden History 141

tree species arriving “naturally” in Western Europe about 2,800 years ago. At this
time, human influence on vegetation had become prominent since the Neolithic,
ca. 7,000 years ago (e.g. Iglesias et al. 2019). Human activity impacts on woodland
included forests clearance for the establishment of cropland and pasture, forest
exploitation for fuel and timber and proto industry, for example, metalworking. New
species importation, such as walnut and chestnut, due to contacts with the
Mediterranean world during the Antiquity (Zohary and Hopf 2000), more than
2,000 years ago, also modified woodlands composition.

In recent decades, many local to regional analyses have shed light on woodland
evolution at the rhythm of socio-ecosystem construction in Western Europe. These
recent palynological studies took place in a framework of multiproxy approaches on
cores such as sediments, micro-charcoals, plants macro-remains analyses and,
recently, on ancient DNA. Cores dating are increasingly precise with many
radiocarbon dates and the use of radioelements on the uppermost part of the cores
(210Pb and 137Cs) (Appleby 2001). Another challenging target of recent studies is the
understanding of past biodiversity and its long-term dynamics using pollen diversity
(e.g. Colombaroli and Tinner 2013).

Figure 11.2. Holocene forest evolution in lowland temperate Europe: the example of
Premery sequence (340 m. asl), Burgundy, France. The main tree taxa pollens are
expressed here as percentages of total land pollen; anthropogenic indicators curve
gathered pollen of crops, weeds and ruderal species (modified from Jouffroy-Bapicot
2014). For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
142 Historical Ecology

Many studies aim to over cross pollen analysis limits related to pollen
identification that rarely reached the species level, and pollen taxa rates that cannot
be directly transformed into land cover (e.g. Seppä 2007). The development of
models and their application on databases such as the EPD
(http://europeanpollendatabase.net/index.php) specifically focus on the development
of quantified and spatially explicit approach of the data (e.g. Marquer et al. 2017).
At a European scale, Roberts et al. (2018) provided a recent synthesis of these three
main different approaches, by transforming fossil data into quantitative past
European forests cover over the last 11,000 years.

11.6. From the air: LiDAR

LiDAR is a remote sensing technique providing highly accurate 3D models of


physical bodies with mobile laser beams. Archaeologists have used this technique
since the early 2000s (Devereux et al. 2005). This data offers an unprecedented
opportunity to the detection, studying and characterization of past land-use in current
forests. Artificially stripped from plant cover, grounds testify of past anthropogenic
activities, still visible at the surface as microreliefs and reveal an amazingly diverse
and abundant heritage which is preserved in forests over past centuries and
millennia. Traces of anthropogenic activities, when visible as microreliefs detectable
with LiDAR, are the remains of buildings, roads, fortifications, mines, quarries,
kilns, field limits, etc., ruined but still existing, sometimes after several millennia of
abandonment. Forests have better preserved ancient structures than cultivated soils
because archaeological remains were erased from landscapes long ago in open areas,
eroded by plowing or destroyed by urbanization. Remains preserved in today’s
forests are of great interest to archaeologists because they are irreplaceable
testimonies of past societies. Past anthropogenic activities have also left traces
undetectable with LiDAR, like buried structures invisible to the surface, or
physicochemical properties of the soil caused, for example, by stone removals for
agricultural purpose, or soil manuring.

In France, 31% of the metropolitan area is covered by woodland today, nearly as


much as arable land. Only a small part of French forests is currently documented
with LiDAR, and an even smaller part has been archaeologically studied.
Investigated forests reveal many traces of past activities of various forms like the
frequent remains of ancient field systems (Figure 11.3). They testify of agricultural
or pastoral activities over the last millennia, often dated from the Roman period (e.g.
Georges-Leroy et al. 2012; Fruchart 2014). Our forests also shelter many other
forms of remains (e.g. Fruchart 2014; Costa et al. 2020): burial sites and buildings,
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Towards a Methodological Framework for Investigating the Hidden History 143

hillforts, military sites, hundreds of ancient road sections, water management


structures, mining remains, quarries, charcoal kilns platforms and limekilns, etc.

Figure 11.3. Comparison between past and present landscapes: the Forêt de
Chailluz in Besancon, France. (A) Today. (B) Hypothetical rendering for the Roman
period. For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip

Archaeological remains are discovered all over the world in today’s forests,
covering large areas, with unsuspected extents and sometimes in unforeseen
locations (e.g. Opitz and Cowley 2013). This new and abundant information,
provided thanks to LiDAR, underlines that we still have little knowledge of the ways
ancient societies used their lands – and forests are particularly poorly known from
this viewpoint. This field of research needs to be developed through interdisciplinary
approaches, to characterize the extents, the patterns and practices associated with
past land-use, and to better understand how anthropogenic activities, combined with
natural dynamics, contributed at landscapes and environment change over time.

11.7. Discussion

As far as the stability of the two major mechanisms causing ecological niche
differentiation over tree life cycle is concerned, i.e. ontogenetic niche shift and
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
144 Historical Ecology

environmental changing conditions during the Holocene period, these five


disciplines and their toolboxes, when crossed used, allow a better analysis,
understanding and interpretation of the effects of anthropogenic activities on
patterns and processes of forests history (Table 11.1). They provide valuable
complementary information that foster a holistic approach to the analysis and
interpretation of evolutionary processes in historical ecology.

Ecological research on woodlands is now widespread all over Europe. Historical


archives offer more variable information that strongly depends on the type of
available documents as well as on the local or regional context. Specialists of forest
history often work with incomplete sets of information alternating large gaps and
areas of abundant data. The comparison and cross-referencing of heterogeneous
documents allow a better reconstruction of the various uses of a forest and a better
visibility of forest developments and dynamics. Likewise, the possibilities of
dendrochronology and palynology depend on the nature and context of
archaeological sites or sedimentary deposits.

Before
12,000 ----- 2000 1000 0
Present
From… Spatial scale
Site
The "surface": Ecology Stand
Landscape
Site
The "wall": History Stand
Landscape
Site
The "wood": Dendrochronology Stand
Landscape
Site
The "ground": Palynology Stand
Landscape
Site
The "air": LiDAR Stand
Landscape

Table 11.1. Fitting temporal and spatial scale according to disciplines


in forest historical ecology. Blank = not suited; gray = weakly suited;
hatched = moderately suited; black = highly suited
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Towards a Methodological Framework for Investigating the Hidden History 145

Nevertheless, several studies using this type of methods have already been
carried out in France in order to answer ecological questions or to focus on
historical, archaeological or paleoenvironmental topics. Some major publications
emphasize the importance of interdisciplinary research that includes ecology,
history, archaeology and paleoenvironmental sciences (Bergès and Dupouey 2017).
Some projects that were developed in varied geographical areas and contexts are
mentioned in Table 11.2.

Study areas Disciplines involved Research topics


Northern France Lowlands Pedology, Understand the dynamics of forest
(alluvial plains) anthracology, and human relationships in a
archaeology, history, temperate deciduous forest since the
LiDAR, Roman period.
for example, Feiss et al.
(2017)
North-eastern France Ecology, pedology, Characterize land-use change since
Lorrain plateau (sandstone, palynology, the Roman period and evaluate
limestone) dendrochronology, environmental consequences and
Jura low mountains and anthracology, legacies to vegetation.
Burgundy plateau (karstic archaeology, history,
limestone) LiDAR,
for example, Dupin
et al. (2017)
Massif Central Ecology, archaeology, Research on land cover change,
Natural Regional Parks history, LiDAR, ancient forests and associated
(crystalline rocks, for example, Renaux biodiversity.
limestone) and Villemey (2016)
South-eastern France Ecology, Evolution of forest management and
Forest of Sainte-Baume, paleoenvironmental of oak (Quercus pubescens) and
Natural Regional Park of approaches, beech (Fagus sylvatica) presence
Luberon (limestone) archaeology, history, over several centuries.
for example, Abadie Land-use legacies in current
et al. (2021) understory vegetation; analysis of
historic archives contribution to the
ecological investigation.
South-western France, Ecology, forestry, Multiple interdisciplinary studies
Pyrenees (crystalline rocks, pedology, since the 1980s, focused on the
limestone) dendrochronology, history and evolution of woodlands,
anthracology, under different headings.
palynology,
archaeology, history,
for example, Saulnier
et al. (2020)

Table 11.2. French examples of studies involved


interdisciplinary in forest historical ecology
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
146 Historical Ecology

This spatial and quantitative reconstruction is necessary for optimizing the


cross-correlation with other data from archaeology and historical archives. The
complexity of datasets often demonstrated a continuity of land-uses at the landscape
level. Paleoecological insights on the evolution from natural to cultural woodland
provide invaluable information for the present and future forest management
(Conedera et al. 2017). An increasing number of tools and datasets now available
including data mining, increasingly accurate dating for paleoecological data and
accessible statistical tools could help achieve the matching of spatial and
chronological scales of the different scientific fields.

11.8. References

Abadie, J., Dupouey, J.-L., Salvaudon, A., Gachet, S. (2021). Historical ecology of
Mediterranean forests: Land use legacies on current understorey plants differ with time
since abandonment and former agricultural use. J Veg Sci, 32(1), e12860.
Altman, J., Hédl, R., Szabó, P., Mazůrek, P., Riedl, V., Müllerová, J., Kopecký, M., Doležal,
J. (2013). Tree-rings mirror management legacy: Dramatic response of standard oaks to
past coppicing in Central Europe. PLoS One, 8(2), e55770.
Appleby, P.G. (2001). Chronostratigraphic techniques in recent sediments. In Tracking
Environmental Change Using Lake Sediments, Last, W.M. and Smol, J.P. (eds). Kluwer
Academic Publishers, Dordrecht.
Badeau, V., Becker, M., Bert, D., Dupouey, J.L., Lebourgeois, F., Picard, J.-F. (1996).
Long-term growth trends of trees: Ten years of dendrochronological studies in France. In
Growth Trends in European Forests, Spiecker, H., Mielikäinen, K., Köhl, M.,
Skovsgaard, J.P. (eds). Springer, Berlin, Heidelberg.
Beck, C. (2008). Les eaux et forêts en Bourgogne ducale (vers 1350 – vers 1480) : société et
biodiversité. Éditions L’Harmattan, Paris.
Bergès, L. and Dupouey, J.-L. (2017). Écologie historique et ancienneté de l’état boisé :
concepts, avancées et perspectives de la recherche. Rev For Fr, 69, 297–318.
Billamboz, A. (2014). Regional patterns of settlement and woodland developments:
Dendroarchaeology in the Neolithic pile-dwellings on Lake Constance (Germany).
Holocene, 24, 1278–1287.
Büntgen, U., Tegel, W., Nicolussi, K., McCormick, M., Frank, D., Trouet, V., Kaplan, J.O.,
Herzig, F., Heussner, K.-U., Wanner, H. et al. (2011). 2500 years of European climate
variability and human susceptibility. Sci., 331, 578–582.
Colombaroli, D. and Tinner, W. (2013). Determining the long-term changes in biodiversity
and provisioning services along a transect from Central Europe to the Mediterranean.
Holocene, 23, 1625–1634.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Towards a Methodological Framework for Investigating the Hidden History 147

Conedera, M., Colombaroli, D., Tinner, W., Krebs, P., Whitlock, C. (2017). Insights about
past forest dynamics as a tool for present and future forest management in Switzerland.
For Ecol Manage, 388, 100–112.
Corvol, A. (ed.) (1999). Les sources de l’histoire de l’environnement, le XIXe siècle. Éditions
L’Harmattan, Paris.
Costa, L., Laüt, L., Petit, C. (2020). Archéologie, forêt et LiDAR : une recherche qui a du
relief. Archéologies numériques, 4(1), 1–7.
Devereux, B.J., Amable, G.S., Crow, P., Cliff, A.D. (2005). The potential of airborne lidar for
detection of archaeological features under woodland canopies. Antiquity, 79(305),
648–660.
Dufraisse, A., Coubray, S., Girardclos, O., Nocus, N., Lemoine, M., Dupouey, J.-L.,
Marguerie, D. (2018). Anthraco-typology as a key approach to past firewood exploitation
and woodland management reconstructions. Dendrological reference dataset modelling
with dendro-anthracological tools. Quat Int, 463, 232–249.
Dupin, A., Girardclos, O., Fruchart, C., Laplaige, C. (2017). Anthracology of charcoal kilns in
the forest of Chailluz (France) as a tool to understand Franche-Comte forestry from the
mid-15th to the early 20th century AD. Quat Int, 458, 200–213.
Feiss, T., Horen, H., Brasseur, B., Buridant, J. (2017). Historical ecology of lowland forests:
Does pedoanthracology support historical and archaeological data? Quat Int, 457, 99–112.
Fruchart, C. (2014). Analyse spatiale et temporelle des paysages de la forêt de Chailluz
(Besançon, Doubs) de l’antiquité à nos jours. PhD Thesis, University of Franche-Comte,
Besançon.
Georges-Leroy, M., Bock, J., Dambrine, É., Dupouey, J.-L. (2012). Les vestiges
gallo-romains conservés dans le massif forestier de Haye (Meurthe-et-Moselle). Leur
apport à l’étude de l’espace agraire. In Des hommes aux champs. Pour une archéologie
des espaces ruraux du Néolithique au Moyen Âge, Carpentier, V. and Marcigny, C. (eds).
Presses universitaires de Rennes, Rennes.
Girardclos, O. and Perrault, C. (2019). Les forêts de chêne du centre-est de la France. In La
forêt au Moyen Âge, Richard, H. and Bepoix, S. (eds). Editions Les Belles Lettres, Paris.
Haneca, K., Wazny, T., Van Acker, J., Beeckman, H. (2005). Provenancing Baltic timber
from art historical objects: Success and limitations. J Archaeolog Sci, 32, 261–271.
Iglesias, V., Vannière, B., Jouffroy-Bapicot, I. (2019). Emergence and evolution of
anthropogenic landscapes in the western Mediterranean and adjacent atlantic regions.
Fire, 2(4), 53.
Jouffroy-Bapicot, I. (2014). New palaeoecological data provided by the Premery forest closed
depressions (Burgundy-France). Quater, 25(3), 253–269.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
148 Historical Ecology

Ludemann, T., Michiels, H.-G., Nalken, W. (2004). Spatial patterns of past wood
exploitation, natural wood supply and growth conditions: Indications of natural tree
species distribution by anthracological studies of charcoal-burning remains. Eur J For
Res, 123, 283–292.
Marquer, L., Gaillard, M.-J., Sugita, S. (2017). Quantifying the effects of land use and climate
on Holocene vegetation in Europe. Quat Sci Rev, 171, 20–37.
Opitz, R.S. and Cowley, D.C. (eds) (2013). Interpreting Archaeological Topography.
Airborne Laser Scanning, 3D Data and Ground Observation. Oxbow Books, Oxford.
Petrequin, P., Arbogast, R.M., Bourquin-Mignot, C., Lavier, C., Viellet, A. (1998).
Demographic growth, environmental changes and technical adaptations: Responses of an
agricultural community from the 32nd to the 30th centuries BC (Chalain, Clairvaux,
archaeology). W Archaeolog, 30, 181–192.
Plue, J., Verheyen, K., Van Calster, H., Marage, D., Thompson, K., Kalamees, R.,
Jankowska-Blaszczuk, M., Bossuyt, B., Hermy, M. (2010). Seed banks of temperate
deciduous forests during secondary succession. J Veg Sci, 21(5), 965–978.
Poublanc, S. (2015). Compter les arbres : une histoire des forêts méridionales à l’époque
moderne. PhD Thesis, University of Toulouse-Jean Jaurès, Toulouse.
Renaux, B. and Villemey, A. (2016). Identifier et caractériser les forêts anciennes du Massif
central. État des connaissances – boîte-à-outils – perspectives. Report, Conservatoire
botanique national Massif-Central, Chavaniac-Lafayette.
Roberts, N., Fyfe, R.M., Woodbridge, J., Gaillard, M.-J., Davis, B.A.S., Kaplan, J.O.,
Marquer, L., Mazier, F., Nielsen, A.B., Sugita, S. et al. (2018). Europe’s lost forests: A
pollen-based synthesis for the last 11,000 years. Sci Rep, 8, 716.
Rochel, X. (2017). Une biogéographie historique. Forêts et industries dans le comté de Bitche
au XVIIIe siècle. Histoire et Mesure, 32(2), 9–38.
Salomón, R., Rodríguez-Calcerrada, J., Zafra, E., Morales-Molino, C., Rodríguez-García, A.,
González-Doncel, I., Oleksyn, J., Zytkowiak, R., López, R., Miranda, J.C. et al. (2016).
Unearthing the roots of degradation of Quercus pyrenaica coppices: A root-to-shoot
imbalance caused by historical management? For Ecol Manage, 363, 200–211.
Saulnier, M., Cunill Artigas, R., Foumou, L.F. (2020). A study of late Holocene local
vegetation dynamics and responses to land use changes in an ancient charcoal making
woodland in the central Pyrenees (Ariège, France), using pedoanthracology. Veget Hist
Archaeobot, 29, 241–258.
Seppä, H. (2007). Pollen analysis, principles. In Encyclopedia of Quaternary Science, Elias,
S.A. (ed.). Elsevier Science, London.
Van Steenis, C. (1956). Basic principles of rain forest sociology. Study of Tropical
Vegetation. Proceedings of the Kandy Symposium, Ceylon. UNESCO, Paris.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Towards a Methodological Framework for Investigating the Hidden History 149

Watt, A.S. (1947). Pattern and process in the plant community. J Ecol, 35, 1–22.
Whittaker, R. (1953). A consideration of climax theory: The climax as a population and
pattern. Ecol Monogr, 23(1), 41–71.
Zohary, D. and Hopf, M. (2000). Domestication of Plants in the Old World: The Origin and
Spread of Cultivated Plants in West Asia, Europe, and the Nile Valley, 3rd edition.
Oxford University Press, Oxford.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
12

The Gate to the Forest


is in its History
Keith J. KIRBY
University of Oxford, UK

12.1. Introduction

Trees, woodlands and forests can only be understood properly through their
history. Individual trees may be several hundred years old; the growth rates shown
in their annual rings reflect past changes in climate and levels of competition from
nearby trees. The forms of trees often point to past management practices such as
coppicing; earthworks such as boundary banks further emphasize previous human
activity. The distribution of trees and woods in the landscape tells of land
clearances, farm abandonments and forest creation.

Different approaches may be used to explore woodland history (Table 12.1;


Watkins 2015). Changes in tree cover over thousands of years can be charted from
pollen grains found at different depths in a soil or peat profile, patterns of fire from
charcoal remains, and shifts in landscape openness from the species composition of
subfossil beetles and snails. Archaeologists interpret the nature of woodland around
settlements by analyzing the wood used in buildings, boats or for firewood. Woods
were valuable properties and wood as a material was widely traded: historic
documents may refer to trees as boundary markers; court reports tell how the
ownerships were challenged and disputes over damage to the trees settled.
Throughout the last 1,000 years, there have been exhortations and instructions as to
how trees and woods should be managed from royal decrees to estate forestry
records, albeit (as now) these were not always followed in practice. Early maps

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
152 Historical Ecology

distinguished wooded from non-wooded lands because of the value of woodland, or


the threat its presence might pose in harboring outlaws and enemy troops. This
chapter explores how such knowledge from the past has also been used to improve
woodland conservation.
Years BP Source and information contained
10,000-2000 Tree composition from peat core (Hone et al. 2001)
c.1700 Roman pottery suggests farming in Wytham area
1070 Anglo-Saxon charter refers to woodland in Wytham area
900 Monastic records for woodland in Wytham area with hunting and firewood rights
c.500 Boundary banks (still recognisable) established around named coppices
The oldest tree surviving in the area, the Broad Oak, starts to grow
480 Accounts of disposal of monastic land refers to named coppices in the Wytham area
400 Vice-chancellor’s account books refer to Wytham Woods, as source of oaks bought by the University
260 First detailed map showing woodland boundaries (further maps through to present day)
Accounts of wood and timber sales in local trade magazine
200 Estate papers describe major plantings and reorganisation of woods
Ring counts show oaks surviving from this period
100 Aerial photograph shows a corner of the Woods with World War I training trenches
80 First student management plan written as part of forestry degree course
70 First complete aerial photographic cover for the Woods
Charles Elton starts to record visits to Wytham in his diaries (Elton 1942-1965)
First formal management plan produced by the University Forestry Department
65 Grayson and Jones produce their history of the Woods.

Table 12.1. Types of evidence used to explore the history of Wytham Woods, near
Oxford in southern England (Grayson and Jones 1955; Savill et al. 2010)

12.2. The ancient woodland idea

A landowner in eastern England speculated that the presence of the bluebell


Hyacinthoides non-scripta might indicate remnants of the “aboriginal” woodland
cover (Beevor 1925). E.W. Jones, an ecologist in the University of Oxford,
commented that the herb paris Paris quadrifolia was in those parts of Wytham
Woods marked on a 1761 map, but not in adjacent plantations created on farmland
in the early 19th century. In the 1970s, extensive historical and field work in
Lincolnshire and East Anglia confirmed that some vascular plants were much more
likely to be abundant in woodland patches that had existed for several hundred years
or more (Peterken 1974; Rackham 2003). A similar association exists across much
of Europe (Honnay et al. 1999; Hermy 2015).

Older woods tend to contain a wide range of interesting vascular plants and are
also more likely to have other associated biological and cultural values (Peterken
1983). In Britain, the term “ancient woods” was adopted for sites that appeared to
have a continuity of woodland cover for the last few hundred years, if not longer. A
project was set up to list these across Britain using historical maps, documents and
aerial photographs (Spencer and Kirby 1992; Goldberg et al. 2007, 2011;
Figure 12.1). This produced a provisional inventory of ancient woods that has
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Gate to the Forest is in its History 153

proved useful in prioritizing action both with respect to local site management and at
a national level. Subsequent changes in forestry and urban policies mean that fewer
ancient woods are now converted to conifer plantations or cleared for housing.

Figure 12.1. Woodland shown on a map (c. 1:10,000 scale) from southern
England (dated AD 1777); most of these patches are still present

12.3. Legacies of woodland management

Rackham (1990, first published 1976) alerted the conservation sector to the
importance of woodland management for maintaining features and species of
interest. His work and studies inspired by him show how estate records and field
work can be combined to reconstruct past woodland composition and structures
(Figure 12.2). Oral history is also important in capturing ordinary, everyday
practices. In Britain, just after World War II, many of the old crafts were dying out,
but it was still possible to record the experience of the workers concerned (Edlin
1949). Grigson (1975) working slightly later (c. 1958) compiled lists of local plant
names and their folk uses, which would not be possible today. Similarly, Bürgi and
Gimmi (2007) used oral history sources to document the often-overlooked practice
of litter collection. Recording the detail of past management is necessary if their
effects are to be understood.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
154 Historical Ecology

Figure 12.2. Boundary bank and old coppice stools of Carpinus betulus for
one of the woods in Figure 12.1. For a color version of this figure,
see www.iste.co.uk/decocq/ecology.zip

Despite increased protection for ancient woodland, many woodland species have
continued to decline even in areas where the extent of ancient woodland remains
high. Changes in historical woodland management, in particular the decline of
coppicing, often underlie these losses of species that depend on open conditions or
young growth stands (Buckley and Mills 2015a). In Britain, species affected include
the butterfly Boloria selene, the bird Luscinia megarhynchos and the woodland plant
Carex depauperata. Open conditions are less widespread under modern high forest
management, because coppice rotations are much shorter than those for high forest
(Buckley and Mills 2015b). Open woodland would also have allowed more frequent
regeneration of oak, which is often only now present as large trees (Chudomelová
et al. 2017). Less obvious changes in the ground flora may be happening as a result
of the decline in collection of litter with the associated removal of nutrients from the
woodland system (Dzwonko and Gawroński 2002).

12.4. Seeing the trees, not the woods

Foresters and woodland ecologists focus on woods and forests, but landscapes
with scattered open-grown trees are common and were often more widespread in the
past. Wood-pastures, where the ground between the trees was kept open by grazing
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Gate to the Forest is in its History 155

(Plieninger et al. 2015), are important for their vegetation mosaics and for large old
trees, important for their saproxylic invertebrates, fungal populations and epiphytes
(Hartel et al. 2015; Sebek et al. 2016). Their history can be more difficult to trace
than that of closed woodland because open woodland might be mapped as closed
woodland (if the trees were dense enough) or as open treeless ground in other
circumstances. However, such sites were often associated with historic royal hunting
areas and with parkland around castles or other large houses (Jonsell 2012; Farjon
2017). These are more likely to be well-represented in written records and appear in
drawings and pictures, perhaps as background to the house or castle.

Individual trees may be recorded as geographic markers, as in Anglo-Saxon


charters (Hooke 2010), Scandinavian path markers (Ericsson et al. 2003) or the
witness trees used in land surveys by European settlers in North America (Whitney
1994). Ancient trees may acquire their own stories and traditions that can generate
support for conservation interest and action in the real landscape history. For
example, the Major Oak in Sherwood Forest attracts visitors because of its supposed
association with Robin Hood. Other trees come to act as local or national symbols,
such the Gernika oak that stood in front of the parliament of the Biscay Province in
the Basque country (Leroy et al. 2020).

In parts of Europe, trees and shrubs themselves form the boundary of fields, from
remnants left after woodland clearance to new lines planted or self-sown along
fences. For the last 200 years, these boundaries have generally been shown on
large-scale maps. Their creation or clearance in response to changing farming
practice has a strong influence on the ecology of the intervening farmland. The
hedge forms a habitat in its own right and may act as a corridor for species
movement through the landscape (Watt and Buckley 1994; Davies and Pullin 2007).

12.5. Exploring the distant past

Human activity has been the dominant force shaping the distribution and
character of trees and woods across much of Europe for the last few thousand years,
but what were the more natural patterns that developed prior to the expansion of
Neolithic farming? Can such landscapes be the baseline model for conservation
(Hodder et al. 2009) or act as a guide for what might develop under future
“rewilding” (Pettorelli et al. 2019)? Much of the potential natural vegetation of
northwest Europe has been described as different types of largely closed forest
(Bohn et al. 2000). However, an alternative view is that the landscape was quite
open and more like a wood-pasture (Rackham 1998; Vera, 2000): this openness was
maintained by populations of large, wild herbivores, particularly the wild ox Bos
primigenius and the bison Bison bonasus (Table 12.2).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
156 Historical Ecology

Predominantly wooded landscape, cf. Peterken Half-open wood-pasture type landscape cf. Vera
(1996) (2000)

Temporary and permanent open space present in a Half-open landscape with much scrub, scattered trees
matrix of closed canopy stands. Openness and stands of trees, ‘groves’ which might be up to tens
probably less than 20% of the total area. or a few hundreds of hectares across.

Closed canopy areas tended to maintain Very limited regeneration within stands, which
themselves by regeneration in gaps within the eventually broke down to open ground; regeneration
stand. to scrub and tree cover concentrated in the currently
open stands.

Large herbivores present, but not the main shapers Large herbivores the main drivers of landscape
of the landscape structure. Regeneration patterns structure and composition. Other drivers much less
generally driven by other factors (wind, disease, significant.
fire, flood, and early human activity, etc).

Shade tolerant trees such as beech and small- Oak remained abundant; succession to more shade-
leaved lime became abundant; oak remained tolerant species was limited by high grazing pressure.
present, but a more minor component.

Development of open managed woodland Abundance of oaks maintained by grazing; shift to


represents a shift from the ‘natural state’. managed woodland and wood-pasture maintained an
analogue of the natural landscape.

Rewilding might be expected to lead to an increase Rewilding should include sufficient large herbivores
in tree cover, tending towards largely-closed to produce shifting mosaics of open, scrub and closed
landscapes grove vegetation.

Table 12.2. Contrasting visions for European natural landscapes in the mid-Holocene

Palynological studies favor a more closed landscape with little evidence for the
extensive grassland or heathland appearing in the pollen record (Mitchell 2005).
Where there is an increase in indicators of open landscapes, this is generally
associated with evidence of climatic change or human intervention (Nielsen et al.
2012; Fyfe et al. 2013).

Analysis of subfossil beetle remains provides evidence for a more open


landscape: many of the species detected depend currently on old open-grown oaks in
relatively sunny situations (Sebek et al. 2016). However, it is not clear how
important grazing animals were in creating and maintaining open areas (Whitehouse
and Smith 2010; Sandom et al. 2014). Robinson (2014) suggests that something like
Vera’s proposed shifting mosaic may have developed following the clearance and
land abandonment that likely characterized early Neolithic farming. Allen (2017)’s
analysis of snail faunas also suggests a link between open landscapes and high
human use, rather than necessarily the activity of wild herbivores.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Gate to the Forest is in its History 157

An unsolved problem is finding a reliable proxy for the density of large


herbivores. The spores of fungi that live on dung can be extracted from peat and soil
cores, but, as yet, there is not a reliable way of linking spore abundance to animal
numbers (Davies 2019). Nor is it clear that a half-open landscape would necessarily
result from the impacts of large herbivores alone (Kirby 2004).

12.6. Trees and woods from the past to the future

The binary division between ancient and recent woodland has proved useful for
assigning broad conservation priorities, but historical ecology shows that there is
more of a continuum in woodland characteristics. Dupouey et al. (2002) highlight
that legacies of Roman-era clearance may still be detectable in present-day
woodland soils in some French forests. Barnes and Williamson (2015) report ancient
woodland indicators common in some woods that from documentary evidence are
clearly recent. Saproxylic insects, lichens and bryophytes found in or on old trees
show that some aspects of “woodland ecosystems” do not depend on being in closed
woodland at all (Figure 12.3).

Figure 12.3. Many saproxylic invertebrates and epiphytic lichens depend on old
trees, but not on woodland. For a color version of this figure, see
www.iste.co.uk/decocq/ecology.zip
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
158 Historical Ecology

Changing woodland management, or stopping it completely, as in strict forest


reserves and under rewilding has significant impacts on the legacy of species and
assemblages that we inherited (Peterken and Mountford 2017). Moreover, future
climatic and atmospheric conditions will be different from those in which ancient
trees and woods developed. We cannot study the effects of future changes without
knowing how trees and woods arrived at the present (Burgi et al. 2020). We should
use the insights from historical ecology to help conserve the meaning in woods from
the past and to develop new meanings for future woods and treescapes.

12.7. References

Allen, M.J. (2017). The southern English chalklands: Molluscan evidence for the nature of the
post-glacial woodland cover. In Molluscs in Archaeology: Methods, Approaches and
Applications, Allen, M.J. (ed.). Oxbow Books, Oxford, UK.
Barnes, G. and Williamson, T. (2015). Rethinking Ancient Woodland: The Archaeology and
History of Woods in Norfolk. University of Hertfordshire Press, Hatfield, UK.
Beevor, H. (1925). Norfolk woodlands from the evidence of contemporary chronicles. Quart.
J. Forest., 19, 87–110.
Bohn, U., Gollub, G., Hettwer, C. (2000). Map of the Natural Vegetation of Europe. Federal
Agency for Nature Conservation, Bonn, Germany.
Buckley, G.P. and Mills, J. (2015a). Coppice silviculture: From the mesolithic to 21st century.
In Europe’s Changing Woods and Forests: From Wildwood to Managed Landscapes,
Kirby, K.J. and Watkins, C. (eds). CABI, Wallingford, UK.
Buckley, G.P. and Mills, J. (2015b). The flora and fauna of coppice woods: Winners and
losers of active management or neglect. In Europe’s Changing Woods and Forests: From
Wildwood to Managed Landscapes, Kirby, K.J. and Watkins, C. (eds). CABI,
Wallingford, UK.
Bürgi, M. and Gimmi, U. (2007). Three objectives of historical ecology: The case of litter
collecting in Central European forests. Land. Ecol., 22, 77–87.
Bürgi, M., Cevasco, R., Demeter, L., Fescenko, A., Gabellier, N., Marull, J., Ostlund, L.,
Santruchkova, M., Wohlgemuth, T. (2020). Where do we come from? Cultural heritage in
forests and forest management. In How to Balance Forestry and Biodiversity
Conservation. A View Across Europe, Krumm, F., Schuck, A., Rigling, A. (eds). EFI,
WSL, Bern, Switzerland.
Chudomelová, M., Hédl, R., Zouhar, V., Szabó, P. (2017). Open oakwoods facing modern
threats: Will they survive the next fifty years? Biol. Conserv., 210, 163–173.
Davies, A.L. (2019). Dung fungi as an indicator of large herbivore dynamics in peatlands.
Rev. Palaeobot. Palynol., 271, 104–108.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Gate to the Forest is in its History 159

Davies, Z.G. and Pullin, A.S. (2007). Are hedgerows effective corridors between fragments of
woodland habitat? An evidence-based approach. Landsc. Ecol., 22, 333–351.
Dupouey, J.L., Dambrine, E., Laffite, J.D., Moares, C. (2002). Irreversible impact of past land
use on forest soils and biodiversity. Ecology, 83, 2978–2984.
Dzwonko, Z. and Gawroński, S. (2002). Effect of litter removal on species richness and
acidification of a mixed oak-pine woodland. Biol. Conserv., 106, 389–398.
Edlin, H.L. (1949). Woodland Crafts in Britain. Batsford, London, UK.
Elton, C.S. (1942–1965). Elton Archive: Notes on the Wytham Area 1942–1965 [Online].
Available at: http://ora.ox.ac.uk/objects/uuid:89c5e479-6003-45ba-bd78-8a8a12858bf1.
Ericsson, T.S., Ostlund, L., Andersson, R. (2003). Destroying a path to the past – The loss of
culturally scarred trees and change in forest structure along Allmunvagen, in mid-west
boreal Sweden. Silva Fennica, 37, 283–298.
Farjon, A. (2017). Ancient Oaks in the English Landscape. Kew, London.
Fyfe, R.M., Twiddle, C., Sugita, S., Gaillard, M.-J., Barratt, P., Caseldine, C.J., Dodson, J.,
Edwards, K.J., Farrell, M., Froyd, C. et al. (2013). The Holocene vegetation cover of
Britain and Ireland: Overcoming problems of scale and discerning patterns of openness.
Quat. Sci. Rev., 73, 132–148.
Goldberg, E.A., Kirby, K.J., Hall, J.E., Latham, J. (2007). The ancient woodland concept as a
practical conservation tool in Great Britain. J. Nat. Conserv., 15, 109–119.
Goldberg, E.A., Peterken, G.F., Kirby, K.J. (2011). Origin and evolution of the ancient
woodland inventory. B. Wildl., 23, 90–96.
Grayson, A.J. and Jones, E.W. (1955). Notes on the History of the Wytham Estate with
Special Reference to the Woodlands, Imperial Forestry Institute, Oxford, UK.
Grigson, G. (1975). The Englishman’s Flora. Paladin, St Albans, UK.
Hartel, T., Plieninger, T., Varga, A. (2015). Wood-pastures in Europe. In Europe’s Changing
Woods and Forests: From Wildwood to Managed Landscapes, Kirby. K.J. and Watkins,
C. (eds). CABI, Wallingford, UK.
Hermy, M. (2015). Evolution and changes in the understorey of deciduous forests: Lagging
behind drivers of change. In Europe’s Changing Woods and Forests: From Wildwood to
Managed Landscapes, Kirby. K.J. and Watkins, C. (eds). CABI, Wallingford, UK.
Hodder, K.H., Buckland, P.C., Kirby, K.J., Bullock, J.M. (2009). Can the pre-Neolithic
provide suitable models for re-wilding the landscape in Britain? British Wildlife
(supplement), 20, 4–15.
Hone, R., Anderson, D.E., Parker, A.G., Morecroft, M.D. (2001). Holocene vegetation change
at Wytham Woods, Oxfordshire. Quat. Newsletter, 94, 1–15.
Honnay, O., Hermy, M., Coppin, P. (1999). Effects of area, age and diversity of forest patches
in Belgium on plant species richness, and implications for conservation and reforestation.
Biol. Conserv., 87, 73–84.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
160 Historical Ecology

Hooke, D. (2010). Trees in Anglo-Saxon England: Literature, Lore and Landscape. The
Boydell Press, Woodbridge, UK.
Jonsell, M. (2012). Old park trees as habitat for saproxylic beetle species. Biodiv. Conserv.,
21, 619–642.
Kirby, K.J. (2004). A model of a natural wooded landscape in Britain as influenced by large
herbivore activity. Forestry, 77, 405–420.
Leroy, T., Plomion, C., Kremer, A. (2020). Oak symbolism in the light of genomics. New
Phytol., 226, 1012–1017.
Mitchell, F.J.G. (2005). How open were European primeval forests? Hypothesis testing using
palaeoecological data. J. Ecol., 93, 168–177.
Nielsen, A.B., Giesecke, T., Theuerkauf, M., Feeser, I., Behre, K.E., Beug, H.J., Chen, S.H.,
Christiansen, J., Dörfler, W., Endtmann, E. et al. (2012). Quantitative reconstructions of
changes in regional openness in north-central Europe reveal new insights into old
questions. Quat. Sci. Rev., 47, 131–149.
Peterken, G.F. (1974). A method for assessing woodland flora for conservation using
indicator species. Biol. Conserv., 6, 239–245.
Peterken, G.F. (1983). Woodland conservation in Britain. In Conservation in Perspective,
Warren, A. and Goldsmith, F.B. (eds). John Wiley & Sons, Chichester, UK.
Peterken, G.F. (1996). Natural Woodland: Ecology and Conservation in Northern Temperate
Regions. Cambridge University Press, Cambridge, UK.
Peterken, G.F. and Mountford, E.P. (2017). Woodland Development: A Long-term Study of
Lady Park Wood. CABI, Wallingford, UK.
Pettorelli, N., Durant, S.M., Du Toit, J.T. (2019). Rewilding. Cambridge University Press,
Cambridge, UK.
Plieninger, T., Hartel, T., Martín-López, B., Beaufoy, G., Bergmeier, E., Kirby, K., Montero,
M.J., Moreno, G., Oteros-Rozas, E., Van Uytvanck, J. (2015). Wood-pastures of Europe:
Geographic coverage, social–ecological values, conservation management, and policy
implications. Biol. Conserv., 190, 70–79.
Rackham, O. (1990). Trees and Woodland in the British Landscape (Revised Edition). Dent,
London, UK.
Rackham, O. (1998). Savanna in Europe. In The Ecological History of European Forests,
Kirby, K.J. and Watkins, C. (eds). CABI, Wallingford, UK.
Rackham, O. (2003). Ancient Woodland: Its History, Vegetation and Uses in England
(Revised Edition). Castlepoint Press, Dalbeattie, UK.
Robinson, M. (2014). The ecodynamics of clearance in the British Neolithic. Env. Archaeol.,
19, 291–297.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Gate to the Forest is in its History 161

Sandom, C.J., Ejrnæs, R., Hansen, M.D.D., Svenning, J.C. (2014). High herbivore density
associated with vegetation diversity in interglacial ecosystems. Proc. Nat. Acad. Sci. USA,
111, 4162–4167.
Savill, P.S., Perrins, C., Kirby, K.J., Fisher, N. (2010). Wytham Woods, Oxford’s Ecological
Laboratory. Oxford University Press, Oxford, UK.
Sebek, P., Vodka, S., Bogusch, P., Pech, P., Tropek, R., Weiss, M., Zimova, K., Cizek, L.
(2016). Open-grown trees as key habitats for arthropods in temperate woodlands: The
diversity, composition, and conservation value of associated communities. For. Ecol.
Manage., 380, 172–181.
Spencer, J.W. and Kirby, K.J. (1992). An inventory of ancient woodland for England and
Wales. Biol. Conserv., 62, 77–93.
Vera, F.W.M. (2000). Grazing Ecology and Forest History. CABI, Wallingford, UK.
Watkins, C. (2015). Methods and approaches in the study of woodland history. In Europe’s
Changing Woods and Forests: From Wildwood to Managed Landscapes, Kirby, K.J. and
Watkins, C. (eds). CABI, Wallingford, UK.
Watt, T.A. and Buckley, G.P. (1994). Hedgerow Management and Nature Conservation. Wye
College Press, Ashford, UK.
Whitehouse, N.J. and Smith, D. (2010). How fragmented was the British Holocene
wildwood? Perspectives on the “Vera” grazing debate from the fossil beetle record. Quat.
Sci. Rev., 29, 539–553.
Whitney, G.G. (1994). From Coastal Wilderness to Fruited Plain: A History of
Environmental Change in Temperate North America, 1500 to the Present, Cambridge
University Press, Cambridge, UK.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
13

Plant Assemblages and


Ecosystem Functioning, a
Legacy of Long-term
Interactions with Large
Herbivores
Christophe BALTZINGER and Anders MÅRELL
INRAE, UR EFNO, Nogent-sur-Vernisson, France

13.1. Introduction

Large herbivores are primary consumers, intermediates in the trophic web


between primary producers and predators. They lie at the center of a rich network of
direct and indirect interactions, both biotic (as prey and with organisms depending
on similar feeding resources) and abiotic (with different compartments shaping their
home ranges). They intervene in several ecological processes from home range to
landscape to regional scale and thereby affect assembly rules in plant communities
and ecosystem functioning. In this chapter, we will first highlight how large
herbivores shape local plant assemblages, as ecosystem dominant interactors. We
will then unveil how ecologists overcame methodological challenges and made it
possible to monitor the effects of large herbivores at different spatial scales over the
long-term. We also summarize the hypotheses related to how they shaped the
ecosystems during the Late Pleistocene and Holocene, and reveal their role in more

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
164 Historical Ecology

current temporal vegetation dynamics, including legacy effects, at the population


and community levels. In conclusion, we mention passive and active rewilding
dynamics, and the role large herbivores play in species and ecosystem conservation
and in degraded ecosystem restoration in a context of global change.

13.2. Large herbivores are ecosystem dominant interactors

Large herbivores include more than 250 wild, introduced and domestic (i.e.
livestock) species, some of which are ruminants (e.g. bovidae, cervidae) and some of
which are not (e.g. equidae). Their respective feeding strategies also vary along a
gradient from grazers (i.e. grass and roughage eaters) to browsers (i.e. concentrate
selectors) depending on their digestive morphology and physiology. Intermediate
mixed feeders, like the red deer (Cervus elaphus), lie in between, feeding on grass
and roughage as well as on richer items (e.g. leaves, buds). The feeding strategy,
feeding selectivity, body mass and abundance require attention when assessing the
effects that large herbivores have on plant communities.

13.2.1. Large herbivores as ecosystem engineers

Besides their trophic interactions, large herbivores are ecosystem engineers (i.e.
“[…] organisms that directly or indirectly modulate the availability of resources to
other species, by causing physical state changes in biotic or abiotic materials. In
so doing they modify, maintain and create habitats […]”, Jones et al. 1994,
Figure 13.1). Wilby et al. (2001) included trophic interactions in the concept of
ecosystem engineering, because it is sometimes difficult to disentangle trophic
interactions from engineering effects (e.g. wild boar rooting in the soil for bulbs and
rhizomes). Herbivores are involved in intrinsic trophic interactions that can affect
the physical state of the woody understory vegetation (i.e. shrubs and young trees)
and its attributes as habitat for other organisms to hide from predators, provide
thermal comfort, nest or feed. They directly consume selected plants and indirectly
affect neglected ones. They also compete with other herbivores relying on similar
resources, for example, caterpillars, on which birds depend to feed their chicks (i.e.
indirect trophic cascading effects). Large herbivores can also predate non-plant life
forms (e.g. earthworms, insects, eggs; Cocquelet et al. 2019). They also affect soil
characteristics, through physical engineering (deep or surface disturbance) that
affects fluxes in the soil, chemical (e.g. nutrients, water) and physical (e.g. mud,
diaspore1) transport.

1. A spore, seed or other structure that allows plant dispersal.


Figure 13.1. The wisent (Bison bison), a keystone species, involved in diverse ecological ecosystem processes (e.g. fur-
Plant Assemblages, a Legacy of Long-term Interactions with Large Herbivores

epizoochory, trampling, herbivory, nutrient cycling), has been reintroduced back to the Southern Carpathians rewilding area.
Artwork by Jeroen Helmer/ARK Nature www.ark.eu/wisent. For a color version of this figure, see www.iste.co.uk/decocq/
165

ecology.zip

Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
166 Historical Ecology

13.2.2. Large herbivores and plant assemblages

Local plant assemblages are shaped from the plant species belonging to the
regional species pool. These plants must undergo three main filters to integrate the
extant plant community (Lortie et al. 2004). Sequentially, we first have dispersal
(i.e. ability to reach a new environment), then the environmental filter (i.e. local
abiotic conditions plants have to tolerate) and finally, the biotic context (i.e.
interactions with other plants or other types of organisms). Davis et al. (2000) detail
how fluctuating resources in local plant communities can explain how newcomers
find opportunities to enter a community. These opportunities depend on diaspore
arrival, the characteristics of the new plant and the susceptibility of the local
environment to being colonized, i.e. the environment’s ability to provide transient
resources (i.e. released locally and/or added from the outside) for plants to establish
and grow.

Large herbivores are involved in each of these filters. They help direct
long-distance dispersal for many plants via endozoochory and epizoochory
(Baltzinger et al. 2019). Dispersed plants bear specific traits (e.g. diaspore
appendage or high releasing height favor epizoochory); they are usually released in
areas with similar habitat conditions. Endozoochory also influences the build-up of
the soil seed bank (e.g. Jaroszewicz 2013 for wisent). More than just modifying the
local environment, large herbivores drive resource heterogeneity (e.g. nutrients, light
and water) and create microhabitats in relation to their habitat selection patterns.
Nutrient deposition in feces/urine and decomposing carcasses create long-lasting
biogeochemical hotspots (Murray et al. 2013). Small-scale disturbances (e.g. litter
scraping, trampling, feeding-associated digging and uprooting) affect soil physical
structure, chemical composition and mesofauna (Mohr et al. 2005). Selective
herbivory and stem breaking (Olmsted et al. 2021) also modify understory vertical
structure, affecting its ability to filter light and delaying forest succession. This
cascades on organisms like arachnids or birds, which depend on specific vegetation
structures for web-building or nesting. Selective herbivory also mediates plant–plant
interactions via apparent competition. Floral herbivory (Mårell et al. 2009) may also
affect plant–pollinator interactions. Endozoochory influences plant germination
ability and timing (Milotić and Hoffmann 2016). Biotic interactions also include
secondary dispersal by dung beetles, predation on diaspores concentrated in feces by
rodents or decomposition by fungi developing on feces that modulate dispersal
effectiveness.

Though each of these three main filters shapes local plant assemblages, escape
from strong biotic context (i.e. intense chronic herbivory for 20 years) highlighted
the environmental filter acting on forest plant assemblages (Chollet et al. 2021).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Plant Assemblages, a Legacy of Long-term Interactions with Large Herbivores 167

13.3. Long-term effects and methodological changes

13.3.1. Paleoecological records

Reconstructing plant–herbivore interactions in paleoecological times requires


specific approaches (e.g. Bradshaw et al. 2003). One approach is to associate
scattered remains found in similar-aged layers. Radiocarbon dating can provide
accurate estimations for the age of the matrix, or direct dates for bony remains more
than 30,000 years (1,000 years = KY, hereafter) old. Ancient vegetation remains
contain plant macrofossils (i.e. seeds, fruits and leaves) where the plants initially
grew, but pollen, another important indicator, may have been dispersed over long
distances, and should be interpreted at larger spatial scales. Herbivore remains
(bones and teeth) are found more rarely, and their occurrence does not indicate
abundance. Often, herbivore remains are retrieved from the same sediment layers as
vegetation remains. Tooth wear can also help determine what large herbivores
consumed during ancient times. Abundance and identity of dung beetles,
coprophilous fungal spores and pollen records can inform about the guild of large
herbivores and past grazing regimes. Archaeological digs and cave deposits
represent main sites of interest combining archaeozoology, anthracology and
paleoanthropology (Defleur and Desclaux 2019).

13.3.2. Modern data

More recently, several methods have been applied to plant–herbivore


interactions. They generally require the spatial comparison of two different levels of
herbivore abundance. For example, presence and absence could be compared by
experimentally excluding ungulates from a given area (e.g. Bernes et al. 2018) or by
investigating natural refuges (e.g. rocky outcrops, Chollet et al. 2013). Areas hosting
various herbivore densities can also be compared via data on different levels of
hunting pressure (Boulanger et al. 2015), accessibility (Martin and Baltzinger 2002),
eradication attempts, drastic population reductions (Tanentzap et al. 2009) or
management options combining fencing and supplemental feeding (Baltzinger et al.
2016). Scientists have also simulated herbivory by clipping vegetation at different
periods of the year (Laurent et al. 2017). In other cases, herbivore impact has been
simulated by jointly clipping vegetation and releasing nitrogen. Persson et al. (2005)
used this strategy to reflect four levels of moose (Alces alces) populations (0, 0.8,
2.4 or 4 ind.km–²). However, their experimental setup could have been more
complete by considering trampling. Tremblay et al. (2006) also manipulated
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
168 Historical Ecology

herbivore density by introducing a known number of individuals into fenced


compounds with predetermined surface areas.

The types of comparisons described above are static, and assessing temporal
changes in plant–herbivore interactions requires longitudinal approaches, for
example, permanent plots resurveyed at different time scales (from seasons to years
to decades, for example, forestREplot2) for plant communities and herbivory
pressure (e.g. browsing index, abundance assessment, Boulanger et al. 2015).
However, the longer the time period considered, the greater the risk of observer
effect and plot relocation issues. Abundance for herbivores outside exclosures is
difficult to tackle and prone to annual changes (but see Boulanger et al. 2018). One
interesting example is the synchronic studies in the Haida Gwaii archipelago, which
took advantage of delayed colonization patterns (i.e. some islands have never been
colonized, some were colonized 20 years ago and others more than 50 years ago;
Stockton et al. 2005). This colonization history has been unveiled by
dendrochronology, dating the last regeneration of palatable trees and the oldest
fraying scars. The uncolonized islands serve as a reference state, making it possible
to test for resilience. Chronosequences in combination with delayed exclusion
experiments have also been used (e.g. Hidding et al. 2013) to test for legacy effects
(i.e. “the ghost of herbivory past”). Finally, time series of aerial photographs can
help study the effects of herbivores on forest succession.

Furthermore, different vegetation metrics (e.g. abundance in terms of cover


percentage/number of individuals, height, morphology, phenology) have been
usually monitored for various plant species in various treatments to make
comparisons at the species, subgroup (e.g. woody vegetation, forbs, graminoids,
bryophytes) and community levels (e.g. species richness, diversity indices) possible.

13.4. Plant–herbivore interactions over the long-term

13.4.1. Quaternary communities of large herbivores and associated


flora

The Quaternary period includes the Pleistocene (2.588 million years to 11.7 KY
before present, BP) followed by the Holocene lasting until the present. In Europe,
this period presents animal remains of roughly 10–12 large herbivores from the late
Pleistocene until the last interglacial period (128–114 KY BP), including mega
herbivores, like giant deer, auroch, straight-tusked elephants (Bradshaw et al. 2003).

2. https://forestreplot.ugent.be/.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Plant Assemblages, a Legacy of Long-term Interactions with Large Herbivores 169

Habitat types are sometimes inferred from the supposed feeding strategy of these
past herbivores. For instance, the penultimate interglacial period (240 KY BP)
hosted a diverse fauna of grazers indicating savanna-like vegetation, whereas the
latest interglacial period had fewer grazers suggesting more forested habitats. The
early Holocene, just after the latest glaciation, hosted a depleted community of
herbivores, which were smaller (Gill 2014). This transition also marks a shift from
more grazing to more browsing herbivore communities, which are less prone to limit
abundant grassland/savanna-like vegetation, with potential consequences on the
occurrence and intensity of natural fires. Although bone remnants and animal
imprints allow the identification of herbivore species, precise knowledge of
their abundance remains challenging (but see section 13.3.1 for the use of
herbivore-dependent organisms). This makes the understanding of past
plant–herbivore interactions more difficult.

13.4.2. The forest in the early Holocene

There is a rich ongoing debate about what the forest in Northwestern Europe
looked like after the last Ice Age in the early Holocene (Bradshaw et al. 2003). Vera
(2000) proposed the “wood-pasture hypothesis”, which suggests that large
herbivores maintained a mosaic of open habitats, allowing light-demanding trees
(oaks Quercus and hazel Corylus, present in the pollen records) and shrubs to persist
within the forest. Under this hypothesis, large herbivores were the main agents of
the open habitats and unpalatable or less consumed shrubs served as a physical
protection for young, palatable trees during regeneration. However, Mitchell (2005)
found high pollen records for oaks and hazel in Ireland in the absence of large
herbivores, thus invalidating the idea that they played a key role in maintaining these
tree species. Alternatively, the “high forest hypothesis” suggests that the early
Holocene was predominantly composed of closed forests. Paleorecords for various
taxa (e.g. tree pollen, fossils, insects and mollusks) support this hypothesis.
However, we do not know how light-demanding species like oaks and hazel
regenerated. Subsequently, Svenning (2002) suggested that there was probably a
wide range of habitats and that open vegetation was more frequent on floodplains
and infertile soils. Uplands would have been mostly closed forests, with some
openings maintained (but not necessarily created) by large herbivores. It is further
supposed that natural fire regimes, either favored or hampered by herbivores, shaped
the diversity of the habitats in pre-agricultural landscapes. For instance, Waldram
et al. (2008) showed that grazing by the white rhinoceros in rich mesic savannas
controlled vegetation height and facilitated access for other herbivores, thus limiting
the size and intensity of natural fires.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
170 Historical Ecology

13.5. Modern vegetation trajectories driven by large herbivores

13.5.1. Herbivory effects

Bernes et al. (2018) recently performed a systematic review of ungulate


herbivory on understory vegetation in temperate and boreal forests (i) at the
community level (abundance and species richness), (ii) at the sub-group level (the
woody understory – shrubs, seedlings and saplings – graminoids, forbs and
bryophytes) and (iii) at the species level, whenever feasible. They also took into
account the herbivore origin (domestic, introduced and native) and their feeding
strategy (grazer, intermediate mixed feeder and browser).

Bernes et al.’s work points out the negative effects of ungulate herbivory on the
abundance and species richness of woody understory vegetation and on bryophyte
abundance, and the positive effects on forb and bryophyte species richness. Both
abundance and height of most tree saplings tended to be limited by herbivory (e.g.
Quercus, Tsuga), and this occurred regardless of the degree of palatability. A single
forb (Alliaria petiolata), exotic and invasive in North America, benefitted from
ungulate herbivory; whereas the abundance of three species was reduced
(Maianthemum canadense, Calluna vulgaris and Vaccinium vitis-idaea) in their
native range.

Ungulate origin also significantly affected vegetation patterns, both in abundance


and composition. Domestic and introduced ungulates most depleted global
understory abundance, though native ungulates had a noticeable impact on the
woody understory. Only domestic herbivores (all grazers in this review) limited the
abundance of graminoids. If we consider species richness, domestic ungulates
favored global understory species richness, specifically forbs, and to a lesser extent
graminoids, whereas native and introduced ungulates reduced woody understory
species richness. The feeding strategy also had an effect; intermediate mixed feeders
and grazers both reduced understory cover, whereas only grazers favored species
richness.

13.5.2. Temporal trajectories

Bernes et al. (2018) also concluded that the current lack of replicated studies
makes it difficult to assess the temporal effects of different types of herbivores over
the long-term (i.e. six years was the median herbivory exposure length). The higher
the herbivory intensity (i.e. herbivore abundance x exposure length), the lower the
abundance for global understory vegetation, the woody understory and shrubs.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Plant Assemblages, a Legacy of Long-term Interactions with Large Herbivores 171

Sapling species richness also decreased with increased herbivory intensity, while
forb species richness and graminoid abundance increased.

Long-term studies generally report trends in plant community changes (e.g.


Ellenberg indicator values for light and nitrogen; dispersal modes; native status;
Grime CSR strategies) or winner/loser species (i.e. increasing/decreasing in the
frequency of occurrence) and their associated traits. Vild et al. (2017) took
advantage of vegetation resampling inside and outside a game preserve, where
herbivore populations had increased from 5 to 55 ind.km–². They found an increase
in global species richness (see also Boulanger et al. 2018) linked to an increase in
ruderal plants, mostly nitrophilous and light-demanding species, at the expense of
woody and endangered species. Endozoochorous plants were generally favored, but
not epizoo- or anemochorous species. However, dispersal syndromes are not
reliable, as they underestimate zoochorous dispersal by ungulates (Green et al.
2022). Over a 50-year period, Wiegmann and Waller (2006) identified 21 winner
and 21 loser plant species out of a total of 73 species. The winners, 16 natives and
five exotics, were common species, mostly grasses and sedges, all with abiotic
modes of dispersal and pollination, and all browsing-tolerant or browsing-resistant
species. The 21 losers were rarer, deer-sensitive (palatable) species. Globally, the
number of species per surface unit decreased over time, the winner species showing
a more spatially consistent pattern and the loser species being more prone to random
disappearances.

Legacy effects have been revealed in different studies, proving that plant
communities maintain a memory of the previous herbivory events that have shaped
their trajectories. Hidding et al. (2013) showed that only eight years of white-tailed
deer (Odocoileus virginianus, hereafter WTD) herbivory after logging changed the
plant community’s early-successional trajectory. Furthermore, Pendergast et al.
(2016) highlighted a lag time in plant community response following WTD
exclusion: species composition inside exclusion fences started to diverge from the
unfenced control plots only after five years. During a combined plant and herbivory
resurvey of a 30-year monitoring, Boulanger et al. (2015) showed that the plant
community changes were correlated with the gradient of deer herbivory pressure.
However, the magnitude of these changes was lower than expected, again
suggesting a delay in the response of the understory to relaxed deer herbivory
pressure. Royo et al. (2010) tested the resilience of over-browsed understory
vegetation after an experimental culling experiment. Few deer-palatable species
responded quickly to the reduction in WTD populations, but species richness in the
understory did not fully recover; browsing-resistant species continued to dominate,
forming recalcitrant understory layers. In the absence of plant refuges providing a
diaspore pool (Chollet et al. 2013), species that have been reduced to sparse or weak
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
172 Historical Ecology

individuals will barely, or never, recover. Nuttle et al. (2014) also stressed the
long-lasting legacy effects that continued for at least 20 years in the Northeastern
US and concluded that deer population reduction alone was not sufficient for plant
community recovery. Additional measures, such as constraining recalcitrant fern
layers, must be taken. Tanentzap et al. (2009) reached similar conclusions in a
context where introduced red deer populations had been drastically reduced (by
92%) over 40 years. Saatkamp et al. (2021) also evidenced very long-term
sheep-pastoralism legacies in Southeastern France, leading to very high plant
species richness.

13.6. Perspectives, rewilding and ecosystem restoration

Large herbivores have shaped past and present landscapes through a diversity of
ecological processes, among which herbivory is not the least. Currently, a recovery
process of wild ungulates (mostly browsers) is taking place in Europe and
populations will probably reach unprecedented abundance levels. This is being
accompanied by the passive return of their natural predators (e.g. lynx, wolf), thus
restoring natural complexity to some extent in current landscapes. Herbivores, as
prey, will probably see their behavior and population dynamics affected, and
vegetation abundance and diversity will certainly change through cascading effects.
A recent worldwide assessment of the ecological functions fulfilled by
contemporary introduced herbivores shows that they partly restore vanished
ecological functions previously provided by ancient herbivores (before the late
Pleistocene extinctions, Lundgren et al. 2020). Current rewilding programs aim to
bring back populations of grazing animals (e.g. wisent in Romania3, Figure 13.1)
capable of maintaining open habitats and safeguarding the associated faunal and
floral diversity (Garrido et al. 2018). Large herbivores create heterogeneity, not only
in their home ranges but also at the landscape scale, and may be dynamic actors in
restoring degraded habitats (Lundberg and Moberg 2003) through the key interactive
ecological processes bridging plants and animals (e.g. herbivory, diaspore and
nutrient transport, physical engineering). Boulanger et al. (2018) demonstrated that
the presence of large herbivores extended the survival of early successional plants
under forest cover; Rico et al. (2014) showed that rotational shepherding favored the
dispersal of non-specialized diaspores. Large herbivores might not only be of help in
restoring degraded habitats, but they might also improve the conservation of species
and ecosystems over the long-term. However, given the natural dynamics involving
large herbivores, active rewilding programs in human-dominated landscapes must be
coupled with effective monitoring of diverse taxa (e.g. native and exotic flora,
entomofauna and avifauna). This will ensure expected ecological outputs and social

3. https://rewildingeurope.com/areas/southern-carpathians/.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Plant Assemblages, a Legacy of Long-term Interactions with Large Herbivores 173

acceptance requirements are met in order to fully support the current UN decade on
ecosystem restoration 2021–20304.

13.7. References

Baltzinger, M., Mårell, A., Archaux, F., Pérot, T., Leterme, F., Deconchat, M. (2016).
Overabundant ungulates in French Sologne? Increasing red deer and wild boar pressure
may not threaten woodland birds in mature forest stands. Basic and Applied Ecology,
17(6), 552–563.
Baltzinger, C., Karimi, S., Shukla, U. (2019). Plants on the move: Hitch-hiking with ungulates
distributes diaspores across landscapes. Frontiers in Ecology and Evolution, 7(38), 1–19.
doi: 10.3389/fevo.2019.00038.
Bernes, C., Macura, B., Jonsson, B.G., Junninen, K., Müller, J., Sandström, J., Lõhmus, A.,
Macdonald, E. (2018). Manipulating ungulate herbivory in temperate and boreal forests:
Effects on vegetation and invertebrates. A systematic review. Environmental Evidence,
7(13), 1–32.
Boulanger, V., Baltzinger, C., Saïd, S., Ballon, P., Picard, J.-F., Dupouey, J.-L. (2015).
Decreasing deer browsing pressure influenced understory vegetation dynamics over
30 years. Annals of Forest Science, 72(3), 367–378.
Boulanger, V., Dupouey, J.-L., Archaux, F., Badeau, V., Baltzinger, C., Chevalier, R.,
Corcket, E., Dumas, Y., Forgeard, F., Mårell, A. et al. (2018). Ungulates increase forest
plant species richness to the benefit of non-forest specialists. Global Change Biology,
24(2), e485–e495.
Bradshaw, R.H.W., Hannon, G.E., Lister, A.M. (2003). A long-term perspective on
ungulate-vegetation interactions. Forest Ecology and Management, 181(1–2), 267–280.
Chollet, S., Baltzinger, C., Ostermann, L., Saint-André, F., Martin, J.-L. (2013). Importance
for forest plant communities of refuges protecting from deer browsing. Forest Ecology
and Management, 289(1), 470–477.
Chollet, S., Baltzinger, C., Maillard, M., Martin, J.-L. (2021). Deer exclusion unveils abiotic
filtering in forest understorey plant assemblages. Annals of Botany, 123(3), 371–381.
Cocquelet, A., Mårell, A., Bonthoux, S., Baltzinger, C., Archaux, F. (2019). Direct and
indirect effects of ungulates on forest birds’ nesting failure? An experimental test with
artificial nests. Forest Ecology and Management, 437(1), 148–155.
Davis, M.A., Grime, J.P., Thompson, K. (2000). Fluctuating resources in plant communities:
A general theory of invasibility. Journal of Ecology, 88(3), 528–534.
Defleur, A.R. and Desclaux, E. (2019). Impact of the last interglacial climate change on
ecosystems and Neanderthals behavior at Baume Moula-Guercy, Ardèche, France.
Journal of Archaeological Science, 104(April), 114–124.

4. https://www.decadeonrestoration.org/.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
174 Historical Ecology

Garrido, P., Mårell, A., Öckinger, E., Skarin, A., Jansson, A., Thulin, C.-G. (2019).
Experimental rewilding enhances grassland functional composition and pollinator habitat
use. Journal of Applied Ecology, 56(4), 946–955.
Gill, J.L. (2014). Ecological impacts of the late quaternary megaherbivore extinctions. New
Phytologist, 201(4), 1163–1169.
Green, A.J., Baltzinger, C., Lovas-Kiss, Á. (2022). Plant dispersal syndromes are unreliable,
especially for predicting zoochory and long-distance dispersal. Oikos, 2022(2), e08327
[Online]. Available at: https://doi.org/10.1111/oik.08327.
Hidding, B., Tremblay, J.-P., Côté, S.D. (2013). A large herbivore triggers alternative
successional trajectories in the boreal forest. Ecology, 94(12), 2852–2860.
Jaroszewicz, B. (2013). Endozoochory by European bison influences the build-up of the soil
seed bank in subcontinental coniferous forest. European Journal of Forest Research,
132(3), 445–452.
Jones, C.G., Lawton, J.H., Shachak, M. (1994). Organisms as ecosystem engineers. Oikos,
69(3), 373–386.
Laurent, L., Mårell, A., Korboulewsky, N., Saïd, S., Balandier, P. (2017). How does
disturbance affect the intensity and importance of plant competition along resource
gradients? Forest Ecology and Management, 391(May), 239–245.
Lortie, C.J., Brooker, R.W., Choler, P., Kikvidze, Z., Michalet, R., Pugnaire, F.I., Callaway,
R.M. (2004). Rethinking plant community theory. Oikos, 107(2), 433–438.
Lundberg, J. and Moberg, F. (2003). Mobile link organisms and ecosystem functioning:
Implications for ecosystem resilience and management. Ecosystems, 6(1), 87–98.
Lundgren, E.J., Ramp, D., Rowan, J., Middleton, O., Schowanek, S.D., Sanisidro, O., Carroll,
S.P., Davis, M., Sandom, C.J., Svenning, J.-C. et al. (2020). Introduced herbivores restore
late Pleistocene ecological functions. Proceedings of the National Academy of Sciences,
117(14), 7871–7878.
Mårell, A., Archaux, F., Korboulewsky, N. (2009). Floral herbivory of the wood anemone
(Anemone nemorosa L.) by roe deer (Capreolus capreolus L.). Plant Species Biology,
24(3), 209–214.
Martin, J.-L. and Baltzinger, C. (2002). Interaction among deer browsing, hunting, and tree
regeneration. Canadian Journal of Forest Research, 32(7), 1254–1264.
Milotić, T. and Hoffmann, M. (2016). Reduced germination success of temperate grassland
seeds sown in dung: Consequences for post-dispersal seed fate. Plant Biology, 18(6),
1038–1047.
Mitchell, F.J.G. (2005). How open were European primeval forests? Hypothesis testing using
palaeoecological data. Journal of Ecology, 93(1), 168–177.
Mohr, D., Cohnstaedt, L.W., Topp, W. (2005). Wild boar and red deer affect soil nutrients
and soil biota in steep oak stands of the Eifel. Soil Biology and Biochemistry, 37(4),
693–700.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Plant Assemblages, a Legacy of Long-term Interactions with Large Herbivores 175

Murray, B.D., Webster, C.R., Bump, J.K. (2013). Broadening the ecological context of
ungulate–ecosystem interactions: The importance of space, seasonality, and nitrogen.
Ecology, 94(6), 1317–1326.
Nuttle, T., Ristau, T.E., Royo, A.A. (2014). Long-term biological legacies of herbivore
density in a landscape-scale experiment: Forest understoreys reflect past deer density
treatments for at least 20 years. Journal of Ecology, 102(1), 221–228.
Olmsted, C.F., Betras, T., Pasquini, S.C., Destefano, S., Faison, E., Carson, W.P. (2021).
Characteristics of stem-breaking by moose (Alces alces, Cervidae): A case-study and
worldwide review. The Journal of the Torrey Botanical Society, 147(4), 304–315.
Pendergast, T.H., Hanlon, S.M., Long, Z.M., Royo, A.A., Carson, W.P. (2016). The legacy of
deer overabundance: Long-term delays in herbaceous understory recovery. Canadian
Journal of Forest Research, 46(3), 362–369.
Persson, I.-L., Danell, K., Bergström, R. (2005). Different moose densities and accompanied
changes in tree morphology and browse production. Ecological Applications, 15(4),
1296–1305.
Rico, Y., Boehmer, H.J., Wagner, H.H. (2014). Effect of rotational shepherding on
demographic and genetic connectivity of calcareous grassland plants. Conservation
Biology, 28(2), 467–477.
Royo, A.A., Stout, S.L., Decalesta, D.S., Pierson, T.G. (2010). Restoring forest herb
communities through landscape-level deer herd reductions: Is recovery limited by legacy
effects? Biological Conservation, 143(11), 2425–2434.
Saatkamp, A., Henry, F., Dutoit, T. (2021). Romans shape today’s vegetation and soils: Two
millennia of land-use legacy dynamics in Mediterranean grasslands. Ecosystems, 24(5),
1268–1280.
Stockton, S.A., Allombert, S., Gaston, A.J., Martin, J.-L. (2005). A natural experiment on the
effects of high deer densities on the native flora of coastal temperate rain forests.
Biological Conservation, 126(1), 118–128.
Svenning, J.-C. (2002). A review of natural vegetation openness in north-western Europe.
Biological Conservation, 104(2), 133–148.
Tanentzap, A.J., Burrows, L.E., Lee, W.G., Nugent, G., Maxwell, J.M., Coomes, D.A. (2009).
Landscape-level vegetation recovery from herbivory: Progress after four decades of
invasive red deer control. Journal of Applied Ecology, 46(5), 1064–1072.
Tremblay, J.-P., Huot, J., Potvin, F. (2006). Divergent nonlinear responses of the boreal forest
field layer along an experimental gradient of deer densities. Oecologia, 150(1), 78–88.
Vera, F.W.M. (2000). Grazing Ecology and Forest History. CABI Publishing, Wallingford.
Vild, O., Hédl, R., Kopecký, M., Szabó, P., Suchánková, S., Zouhar, V. (2017). The paradox
of long-term ungulate impact: Increase of plant species richness in a temperate forest.
Applied Vegetation Science, 20(2), 282–292.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
176 Historical Ecology

Waldram, M.S., Bond, W.J., Stock, W.D. (2008). Ecological engineering by a mega-grazer:
White rhino impacts on a South African Savanna. Ecosystems, 11(1), 101–112.
Wiegmann, S.M. and Waller, D.M. (2006). Fifty years of change in northern upland forest
understories: Identity and traits of “winner” and “loser” plant species. Biological
Conservation, 129(1), 109–123.
Wilby, A., Shachak, M., Boeken, B. (2001). Integration of ecosystem engineering and trophic
effects of herbivores. Oikos, 92(3), 436–444.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
14

A Historical Ecology of the


Compiègne Forest (N France)
Jérôme BURIDANT, Boris BRASSEUR, Hélène HOREN,
Emilie GALLET-MORON and Guillaume DECOCQ
Jules Verne University of Picardie, Amiens, France

14.1. Introduction

The Compiègne forest (49°22′ N; 2°54′ E; 32–148 m altitude) is the third biggest
state forest in the French lowlands. It is a sub-Atlantic temperate oak-beech forest
located at 70 km north-east to Paris, which covers ca. 14,500 ha at the junction
between the Aisne and the Oise rivers. The substrate is dominated by acidic sands
and limestones. It has long been considered as a very ancient woodland, which
might have continuously existed since the beginning of the last post-glacial
afforestation episode, ca. 10,000 years ago. This belief relies on the fact that (i)
Historians of the 19th century erroneously interpreted Caesar’s landscape
description (De Bello Gallico) as a dense woodland at the time of the Romans
conquest (1st century BC); (ii) following the Roman empire collapse (5th century
AD), the forest became a royal hunting reserve until the French revolution in 1789
and (iii) since the 18th century, the forest is a state forest managed for wood
production and stag hunting. However, evidence is accumulating that the
Compiègne forest may not be as ancient as previously thought, and has been
patterned by thousands of years of human activities. Here, we present a synthesis of
recent research in historical ecology that has been conducted in this forest by the
EDYSAN research group in Amiens (Jules Verne University of Picardie) to better

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
178 Historical Ecology

understand the legacy of past human–forest interactions in the current forest


ecosystem.

14.2. The ancient forest: an intensively managed agricultural


landscape?

Landscape archaeology using aerial photographs developed in England and


France from the 1950s onward (Dassié 1978). In northern France, it revealed a huge
density of archaeological imprints (e.g. buried calcareous stones into silty topsoils)
across agricultural landscapes, mostly of Gallo-Roman origin, indicating that the
land was already intensively cultivated at that time (Agache 1978). In comparison,
woodlands remained largely unexplored since the vegetation cover hides
micro-reliefs, buried walls and other archaeological remains. Extensive archaeology
using walking transects in the understories can help in finding visible topographic
anomalies and artifacts, but may be limited by a dense ground vegetation or by
management-induced disturbances.

In the case of large forests, such as the Compiègne forest, which were already
present during the “forest minimum” (ca. 1850 in France; Bergès and Dupouey
2017) and have been continuously mapped over the last few centuries, it has been
postulated that they likely exist since prehistoric times. This hypothesis was
seemingly supported by ancient written sources. Pliny the Elder and Suetonius, for
example, designed Gaul as Gallia comata, which means “hairy Gaul”, the Gaul prior
to the Roman conquest (Clavel-Lévêque 1989). This was interpreted by historians of
the 19th century as evidence for a land covered by forests (Maury 1848, 1850).
However, the first archaeological explorations of the 19th century started to cast
doubt on the ancient forest hypothesis, since a number of Gallo-Roman sites were
discovered, such as vici (i.e. urban areas), villae (i.e. isolated farms), a sanctuary and
several roads.

From 1997 to 2017, an intensive field survey was implemented, which revealed
more than 300 Roman sites, mostly rural settlements, within the forest (Thuillier
2017). This field survey was complemented by a remote sensing survey using the
LiDAR technology (Box 14.1). A first flight covering a small part of the forest was
conducted in 2011 at 9 points per m²; a second one covered the entire forest in 2014
at a higher resolution (23 points per m²). A total of 5,710 anomalies were evidenced,
including 1,457 one-off (e.g. depressions, bomb holes), 3,797 linear (e.g. banks,
roads) and 456 zonal (e.g. mounds, enclosures) anomalies (David 2014). All
anomalies that cannot be identified with certainty were verified in the field. Some of
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
A Historical Ecology of the Compiègne Forest (N France) 179

them corresponded to prehistoric or protohistoric artifacts (e.g. dolmen, cutting


workshops, barred spurs), but the vast majority of them dated back the Gallo-Roman
period (1st–4th century AC). Only a sandy area located towards the center-east of
the forest was free from occupation sites (Figure 14.1).

LiDAR has rapidly become a major tool in historical ecology. LiDAR (for light
detecting and ranging) is a technique for measuring distances using the properties of laser
beams. The principle is to analyze the return time of a light beam between its emission
source and an object. In forest, topographic LiDAR is mainly used. It is a LiDAR carried
by an aircraft flying at low or medium altitude, measuring the topography over a
relatively large area. It allows us to assess the relative altitude of the ground and objects at
a high precision (a few cm). The laser beam is reflected by the canopies, by tree trunks
and branches and by the ground. It is therefore possible to classify the cloud of points
obtained by selecting the upper points, the intermediate points and the ground points, to
develop a digital elevation model (DEM), a digital terrain model (DTM) and intermediate
models. DTMs have archaeological or geomorphological applications. Elevation models
and intermediate models rather have dendrometric applications, to assess standing timber
volumes and the structure of forest stands. The higher the number of impacts reaching the
ground (in points per m2), the higher the resolution of the resulting model. In Western
Europe, the first archaeological applications of LiDAR were developed in the Rastatt
forest, in Baden (Sittler and Hauger 2007). In France, the first archaeological survey was
carried out in the forest of Haye, east to Nancy (Georges-Leroy et al. 2011). LiDAR is
now used in a variety of applications, and several European countries have developed
programs to cover the entire national territory.

Box 14.1. The LiDAR technology

The results show that during the Gallo-Roman period, human activities were
structured by an important Roman road, connecting two cities: Soissons (Augusta
Suessionum), to the west, and Senlis (Augustomagus), to the south-west. The
primary road crossed three secondary towns (vici), and several farms (villae) were
scattered on both sides, the more distant ones being connected to the main road by
secondary roads. The LiDAR also revealed that over almost the whole forest area,
the land was divided into rectangular parcels, which were also consistently
structured by the ancient road network (Figure 14.2). They are very similar to those
retrieved in other French forests, which have been shown to be the ancient
agricultural lands, where plowing led to large geometric areas surrounded by banks
possibly topped by a hedgerow (Georges-Leroy et al. 2011). In some parts of the
forest, smaller parcels of more irregular shape sit on the top of these typical parcels
of lands, probably of medieval origin (5th–15th century) (Buridant et al. 2020). This
indicates that parts of the current forest were still cultivated in the Middle Ages.
180
Historical Ecology

Figure 14.1. Former Gallo-Roman settlements in the current Compiègne forest. The top-right map shows the location of the
forest in Northern France. The top-left image is a digital terrain model from the LiDAR, which reveals the ancient
parcels of agricultural land. For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip

Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
A Historical Ecology of the Compiègne Forest (N France) 181

Figure 14.2. LiDAR image of part of the Compiègne forest. The current network of
star-shaped roads (1) is inherited from the 16th to 18th century (yellow lines). We can
see under this network the rural territory organization from the Antiquity made up of
large rectangular agricultural parcels (2), a residential area (3) and limestone
extraction areas (4). For a color version of this figure, see www.iste.co.uk/decocq/
ecology.zip

An interesting feature of this “fossilized” agricultural landscape revealed by the


LiDAR-derived digital elevation model is the presence of many closed circular
depressions (CCD) that are systematically located along an inner edge of the
rectangular parcels (Figure 14.3). They consist of 0.5 to 3 m-deep depressions with a
diameter ranging from 20 to 60 m. A pedo-sedimentary analysis of cores taken
outside, on the edge and at the center of 10 CCD, was conducted to determine their
origin. The shallowest CCD (sCCD) revealed a single-phase, non-stratified sediment
infill. The filling material was clearly anthropogenic, consisting of a mixture of local
soils, their limestone substrate and sometimes fragments of ceramics. At the top, soil
horizon development occurred. In contrast, the deepest CCD (dCCD) showed no or
little infilling and the limestone substrate was encountered at <50 cm. Both CCD
types had a bathtub shape with a flat bottom at the interface with the geological
substrate, which is typical of an anthropogenic excavation. However, CCD could not
be limestone quarries used to provide the building material – that are also present
elsewhere in the landscape – since the hard limestone (e5b) was intact. The
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
182 Historical Ecology

chemistry of the local substrate also excluded a karstic (sinkhole) or salt diapir
dissolution origin (Delafosse 1948; Etienne et al. 2011). These CCD thus correspond
to ancient pits, where limestone powder and calcareous marls and loess (e5b/c) were
extracted. As we found no accumulation or overlying of these materials around
CCD, we hypothesize that they were sprayed over agricultural lands to buffer the
natural soil acidification under a humid temperate climate, thereby maintaining
productivity. Liming was indeed a common practice for Gauls and Gallo-Roman, as
mentioned by Latin agronomists (Varro, Pliny the Elder, Columella). The abundance
of CCD (>100) in the Compiègne forest indicates that liming was a widespread
practice during the Roman times. It not only creates an artificial topography and
localized thick anthropogenic soils but also likely induced a general alkalinisation of
soils, which durably impacted soil properties in former fields. Recent work suggests
that liming effects on soil pH last several millennia (Brasseur et al. 2018), and hence
can still impact vegetation and ecosystem functioning (Box 14.2).

To test whether past Gallo-Roman settlement still affect the present characteristics of
soils, vegetation and soil seed bank, we designed a comparative study (Plue et al. 2008).
To each of 24 former isolated Gallo-Roman villae (i.e. “sites”, where an inhabited
building used to be settled), we paired an unoccupied site (i.e. “controls”, free from any
former human settlement but not from past human influence) at a distance of <200 m,
which shared the same topography, substrate and management history. An 800 m2 quadrat
was disposed in each site and control, into which all vascular plant species were recorded,
and soil samples were taken for further soil chemical analyses and permanent seed bank
assessment (25 cores per quadrants, 26 cm diameter, 20 cm depth after litter removal).

The soil was persistently altered on the sites, resulting in elevated phosphorus levels
and pH (dependent on initial soil conditions). The upper soil layers also contained
scattered stone fragments, pieces of limestone and shards of roof tiles (tegulae). These
altered soil features translated into significantly increased vegetation cover and seed bank
density (1,500 seeds m-2 vs. 333 seeds m-2), increased species richness in both vegetation
(37.1 vs. 23.2) and soil seed bank (6 vs. 4) and increased compositional similarity among
sites (Raup & Crick index: 0.53 vs. 0.76), as compared to unoccupied sites. In particular,
the vegetation on Gallo-Roman sites had a significantly more competitive and ruderal but
less stress-tolerant character than controls. Nitrogen-demanding and calciphileous species
were strongly associated with sites, while acidiphilous species were significantly
associated with controls. It is noteworthy that fruit trees (e.g. apple tree, medlar tree) often
appeared in the woody layers on the archaeological site or nearby. The local persistence
of ruderal species below forest canopies is surprising. It is likely facilitated by forest
management, which allows periodic light arrival at the forest floor, hence their
recruitment from the soil seed bank and the replenishment of the latter.

These results revealed that 1,600 years after their reclamation and forest colonization,
Gallo-Roman settlements still leave a persistent trace on forest soil, vegetation and seed
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
A Historical Ecology of the Compiègne Forest (N France) 183

bank. By creating entirely different habitats with irreversibly (?) altered geochemical
features, these archaeological sites form plant species-rich islands in an otherwise
species-poor forest ocean, thereby contributing to biodiversity at the entire forest scale
(Closset-Kopp and Decocq 2015). Furthermore, the Roman legacy in vegetation is not
limited to former residences. Roman roads have likely served as important corridors for
species, before and after their reclamation. For example, in the Compiègne forest, the
archaeophyte Vinca minor has been observed only on the former Roman road.

Box 14.2. Impacts of Roman settlements on biodiversity and ecosystem functioning

The area of the present-day forest of Compiègne was thus previously not a
forest, but rather looked like an open agricultural land. Archaeological studies and
pollen analyses suggest that the local deforestation started long before the Roman
conquest (Perrière and Leroyer 2006). The first populations of farmers settled
around 4500 BC in the Oise and Aisne valleys, which border the current Compiègne
forest at the west and north, respectively (Figure 14.1). At that time, the pollen
diagrams are dominated by oak (Quercus) and hazelnut (Corylus), with lime (Tilia),
elm (Ulmus) and alder (Alnus). This typical vegetation of the Atlantic period
(5500–2300 BC) was probably an open-canopy forest. Beech (Fagus sylvatica) first
appeared after 3000 BC and became common only after 1250 BC (Leroyer et al.
2017).

Figure 14.3. Left: LiDAR-derived digital elevation model revealing closed circular
depressions filled (sCCD) or not (dCCD) by sediments. Other archaeological features
are visible: a Gallo-Roman villa (RV), Senlis-Soissons primary road (PR), a secondary
road (SR) and agricultural parcel limit (APL). Right: Geological profile of a dCCD (top)
and sCCD (bottom). e5b/c Eocene (upper Lutetian) powdery limestone, e5b Eocene
(middle Lutetian) hard limestone. 1, 2 and 3 show the core position for each of the
CCD. For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
184 Historical Ecology

14.3. The Medieval forest: a woodland (re)birth or a savanna-like


ecosystem?

Gallo-Roman settlements were quite suddenly abandoned towards the end of the
4th century AD, with archaeological evidence for destruction and fires as soon as the
3rd century AD (Frémont and Woimant 1975, 1976; Tuffreau-Libre 1977). A
woodland is mentioned for the first time at the 6th century by Grégoire de Tours
(Histoire des Francs), when he relates the death of the Merovingian king Clotaire I
in 561, from a fever “while he was hunting in the forest of Cuise” (dum in Cotia
silva venationem exercet). The forest of Cuise is presented as a hunting forest
belonging to the royal domain. Since this first mention, historical texts confirm the
continuous presence of a woodland in this area, which was later renamed
Compiègne forest.

The forest colonization of abandoned fields occurred in the course of secondary


succession, so that the early Medieval forest may have consisted of relatively dense
oak-dominated woodlands. This hypothesis is supported by the low human
population density in this area, following the Great Invasions, and the fact that the
wood framework of local houses was built using tall trees with very narrow tree
rings, suggesting they grew in packed populations, a typical characteristic of
early-successional forest communities. The human pressure on the forest likely
increased again with the demographic booming of the 12–13th century.
Archaeological studies evidenced activities associated with the supply of wood, such
as pottery workshops (starting at the 10–11th century), glass factories (12th century)
and metalworking industries (14th century), all established into the forest. The
number of glass factories increased in the 14th and 15th centuries, when several
families of glass-makers immigrating from Eastern France established there. At this
time, the bracken fern (Pteridium aquilinum) was likely abundant, since it was the
only source of lye.

Dating of buried charcoals retrieved from kilns in topographic depressions in the


center of the forest suggest that charcoal production was conducted at least from the
12–13th century until the years 1610–1670. However, no written source specifies
the structure, extent and composition of the forest before the 16–17th century, so
that it is challenging to reconstruct the Medieval forest. The latter likely has been
patterned by two main drivers: user rights and hunting.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
185

For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip


A Historical Ecology of the Compiègne Forest (N France)

Figure 14.4. State of play for user rights in the Compiègne forest in 1663.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
186 Historical Ecology

Pedological pits were made in the center of the forest, close to a dune-like relief free
from any archaeological artifact as revealed by the LiDAR imagery. The pits showed
podzols covered a more or less thick layer of drift sands. The abundance of dung beetle
galleries observed in one horizon of the profile (see Figure 14.5 below) suggests locally
abundant animal waste associated with a pasture land-use, and thus a rather open
landscape. Consistently, the sand layer observed on the top of the paleosol may reflect
intensive grazing-induced drift sands on sandy soils.

Figure 14.5. Podzol profile in the drift-sands area (left panel). Close up
of the contact between the dark organo-mineral horizon and the ash grey
eluvial horizon just below (right panel). For a color version of this figure, see
www.iste.co.uk/decocq/ecology.zip

Podzol profile in the drift sands area (left panel). The soil is buried under ca. 35 cm of
sand. The abundance of dung beetle galleries observed at contact between the dark
organo-mineral horizon and the ash gray eluvial horizon just below (right panel) suggests
local soil enrichment associated to a pasture land-use.

To obtain a picture of past local species assemblages, we sampled soil charcoals in a


podzol and two others soil types within currently mature forest stands (Feiss et al. 2017b).
The radiocarbon dating returned no charcoals older than the 10–11th century. We found
no relationship between the charcoal age and their vertical distribution, due to soil mixing
associated with belowground biological activity. Soil charcoal assemblages thus hardly
reflect temporal changes in forest stand composition. However, the fact that the oldest
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
A Historical Ecology of the Compiègne Forest (N France) 187

charcoals came from birch (Betula) supports the hypothesis that the 10–11th-century
forest was very open, with a substantial proportion of shade-intolerant tree species.
Light-demanding, early-successional taxa, while missing from current tree communities,
were almost systematically more abundant than shade-tolerant species across the three
sites. This suggested that the forest canopy used to be much more open than today,
consistent with a savanna-like landscape maintained by the intense human pressure and
heavy grazing by wild and domestic animals (Feiss et al. 2017a).

Box 14.3. What do soil archives teach us?

Like many European forests in the Middle Ages, the Compiègne forest was not
exclusively intended for wood production. Many uses were practiced by the local
populations, such as picking up dead wood, cutting green wood and grazing by
cattle and pigs. From the 13 to 14th century onward, these uses became user rights
granted to local populations and institutions. These user rights were progressively
restricted, especially by the forest reformation of 1663, to limit the overexploitation
of the forest resources (Figure 14.4). Still, such rights were conceded to
16 communities of inhabitants, 25 ecclesiastical communities and one glass furnace.
Some communities have their rights restricted to certain areas, while many others
can take their herds throughout the entire forest. The authorized load often exceeded
one animal per hectare, a density much greater than the carrying capacity of the
forest. It is thus likely that the forest landscape at this time rather looked like a
wooded pasture (i.e. a temperate savanna; Rackham 1998) than a closed-canopy
woodland, a hypothesis supported by soil charcoal analyses (Box 14.3).

At the same time (from the 17th century onward), with the progress in firearms,
pheasant hunting became a fashion. Kings Louis XIV and Louis XV built seven
pheasantries in the center of the forest. Pheasantries are closed, cultivated parks
dedicated to the breeding of pheasants, which were then released in large open
spaces (called “tirés”) into the forest, allowing them to take off. This practice was
still ongoing in the 19th century, since Emperor Napoleon III was a lover of
pheasant hunting and was often coming in the Compiègne forest for this purpose.

However, the Compiègne forest was also dedicated to the production of wood. In
the 17th and 18th centuries, logs were floated to Paris via the Oise river, where they
were used for construction and as fuel. Interestingly, the forest provided a lot of
short, curved logs to royal shipyards on the Atlantic coast, as well as to a local
shipyard, which built barges for inland water shipping. The abundance of short,
curved logs suggests that the core of the forest, which was managed as a high forest
for hunting, was surrounded by more intensively managed short-rotation, dense
coppice-like woodlands.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
188 Historical Ecology

14.4. The contemporary forest (19th century onward): a closed-canopy


multifunctional woodland

Today’s forest landscape is mostly a legacy of recent periods of restoration and


replanting, which started in the mid-18th century (Figure 14.6). From 1772 to 1792,
deciduous trees, mainly pedunculate oaks, were planted by Pierre-Lucien Pannelier
d’Annel and his son, Antoine-Lucien Pannelier. A second phase of reforestation
occurred throughout the 19th century, over 6,610 ha (i.e. 46% of the total forest
area) (Buridant 2008). Following World War II, conifers, mainly Scots pine, were
widely planted over the less fertile areas. More than half of the forest is therefore
composed of recent plantations (<250 years old). It is likely that human disturbances
associated with these plantations have facilitated the invasive spread of the
American black cherry (Prunus serotina), which started towards the mid-19th
century. The Compiègne forest is currently the most invaded in France, with ca.
80% of the forest area where the invader occurs, as isolated trees to monospecific
stands, thereby significantly impacting the recipient ecosystem (Box 14.4).

Figure 14.6. Tree plantation in the Compiègne forest from 1774 to 1974.
For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip

The current forest is mostly dedicated to timber production (oak, beech and Scots
pine) and stag-hunting. As a public sub-urban forest, the forest also experienced an
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
A Historical Ecology of the Compiègne Forest (N France) 189

increasing pressure from local populations for their recreative activities. Two nature
reserves have been created, one in the 19th century (Beaux Monts) and the other in
the 1970s (Grands Monts). These reserves are useful tools to analyze the impact of
the current forest management on vegetation (Box 14.5).

The first decades of the 21st century are marked by rapid changes in the forest
vegetation in response to global changes (Box 14.5). These changes are event
accelerating: most canopy tree species experienced severe dieback due to several
non-independent factors: repeated drought episodes (beech), cockchafer (oak) and
bark beetle (spruce) boom, chalararosis epidemics (ash), etc. As a result, light arrival
at the forest floor increases with dramatic changes in the ground vegetation to come.

Prunus serotina either escaped from the castle park or has been introduced directly in
the Compiègne forest around 1850 but has been reported as an invader only in 1970. It
has now spread in the entire forest. Many studies have been conducted to assess the
impacts of the invasion, by comparing tree physiology, plant communities, soil properties
and disturbance history between P. serotina-invaded and uninvaded paired stands.
Overall, species assemblages in invaded and uninvaded stands were similar (Chabrerie
et al. 2008). Habitat conditions and disturbances explained most of the variation in both
plant diversity and P. serotina density, the last two factors being weakly associated. The
herb layer of invaded stands exhibited significantly more specialist species and a lower
trait diversity compared to uninvaded stands. Traits characterizing shade-tolerant,
short-living ruderals and shade-avoiders (vernal geophytes) were significantly associated
with invaded stands, while those associated with light-demanding graminoids
characterized uninvaded stands. This indicates that P. serotina was becoming the new
ecosystem engineer, first by inducing trait convergence and community specialization,
thus promoting traits that enable species to capture resources in the new environment it
was creating, and second by reducing the grain of local heterogeneities (Chabrerie et al.
2010). Compared to indigenous tree species, P. serotina exhibited higher foliar N, foliar P
and lower foliar C:N and N:P ratios (Aerts et al. 2017). P. serotina also affected foliar
nutrient contents of co-occurring indigenous tree species leading to decreased foliar N and
increased C:N ratio in beech, decreased foliar N:P ratio in hornbeam and beech and
increased foliar P in Scots pine in invaded versus uninvaded stands. P. serotina thus
changes nitrogen, phosphorus and carbon cycles to its own advantage, hereby increasing
carbon turnover via labile litter, affecting the relative nutrient contents in the overstory
leaves and potentially altering the photosynthetic capacity of the other broadleaved
species.

Box 14.4. Impacts of Prunus serotina invasion


on forest vegetation
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
190 Historical Ecology

Recent vegetation changes have been recently documented using a resurvey study and
a life-history trait-based approach (Closset-Kopp et al. 2019). This study highlighted
decadal changes (1970–2015) in the ground vegetation and searched for the drivers
behind these changes. A number of process-based hypotheses were tested using paired
comparisons (old vs. new records) of community-weighted means and conditional
inference classification trees for a number of traits across soil and management types.
Compared to unmanaged stands (nature reserves), managed stands underwent the most
pronounced vegetation changes, as a likely effect of modifications in forest harvesting
practices since 1970. Consistent with general trends observed in European forests, species
richness of vascular plants increased at both stand (α-diversity) and forest (γ-diversity)
scales, due to the non-random, directional colonization by the same suite of species,
causing compositional, functional and phylogenetic homogenization among habitats (i.e.
decreased β-diversity). Forest management, via the repeated passing of heavy forestry
vehicles, emerged as the key driver of these changes. By altering canopy structure, it
increased not only light availability at the forest floor but also the vulnerability of
understories to climate warming and atmospheric deposits. Furthermore, increasingly
heavy forestry vehicles that drive more frequently across forest stands likely generate
microhabitats suitable for ferns, graminoids and N-demanding forbs and also act as
dispersal agents.

Box 14.5. Vegetation changes at the Anthropocene

14.5. Conclusion

Altogether, these results demonstrate that the Compiègne forest has established
on an ancient cultivated plain and on neighboring plateaus, where many people used
to live, work and trade for centuries. In this context, little place was left for
woodlands during the Gallo-Roman period. The rural landscape rather looked like
an intensively managed open field, maybe with a few small woodland patches
restricted to the steepest slopes of surrounding hills and central sand dunes. The
forest established progressively during the early Middle Ages, from the spontaneous
afforestation of abandoned farmlands. Since the secondary succession was strongly
constrained by human activities, wild and domestic herbivores, it is likely that the
vegetation rapidly shifted from a dense early-successional woodland to an open
wooded pasture, so that the late Medieval forest likely resembled more a temperate
savanna than a closed-canopy woodland. Hunting activities have been a major driver
of the forest landscape through medieval, modern and contemporary times, leading
to the long-lasting maintenance of open habitats within a high forest. Canopy
closure and the subsequent development of a shade-tolerant flora is probably a
recent phenomenon, which is currently challenged by global warming-associated
environmental changes since the forest canopy tends to re-open. Most importantly,
we showed that past land-use and historical human activities strongly shaped soil
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
A Historical Ecology of the Compiègne Forest (N France) 191

and vegetation features, thereby influencing today’s ecosystem functioning. Current


biodiversity patterns are thus a composite legacy of each of the historical periods
experienced by the forest, which cannot be understood without a historical ecology
perspective.

14.6. References

Aerts, R., Ewald, M., Nicolas, M., Piat, J., Skowronek, S., Lenoir, J., Hattab, T.,
Garzón-López, C.X., Feilhauer, H., Schmidtlein, S. et al. (2017). Invasion by the alien
tree Prunus serotina alters ecosystem functions in a temperate deciduous forest. Front.
Plant Sci., 8, 179.
Agache, R. (1978). La Somme pré-romaine et romaine. Société des Antiquaires de Picardie,
Amiens.
Bergès, L. and Dupouey, J.L. (2017). Écologie historique et ancienneté de l’état boisé :
concepts, avancées et perspectives de la recherche. Rev. For. Fr., 69, 297–317.
Brasseur, B., Spicher, F., Lenoir, J., Gallet-Moron, E., Buridant, J., Horen, H. (2018). What
deep-soil profiles can teach us on deep-time pH dynamics after land use change? Land
Degr. Dev., 29, 2951–2961.
Buridant, J. (2008). Environnements forestiers, économie et sylviculture dans la France
septentrionale, XVIIe–XIXe siècles. HdR Thesis, Université de Paris IV-Sorbonne, Paris.
Buridant, J., Pichard, C., Gallet-Moron, E. (2020). L’utilisation de la technologie lidar à la
connaissance archéologique et géohistorique : exemples français. In Temi et metodi della
ricerca storica et filologica, nuove riflessioni tra Est ed Ovest, Şipoş, S., Cepraga, D.O.,
Ardelean, L., Cosma, I. (eds). Transylvanian Review, vol. XXIX, suppl. no. 1.
Chabrerie, O., Verheyen, K., Decocq, G. (2008). Disentangling relationships between habitat
conditions, disturbance history, plant diversity and American Black cherry (Prunus
serotina Ehrh.) invasion in a European temperate forest. Div. Distrib., 14, 204–212.
Chabrerie, O., Loinard, J., Perrin, S., Saguez, R., Decocq, G. (2010). Relating understorey
plant species traits to environmental changes in forest stands heavily invaded by Prunus
serotina. Biol. Inv., 12, 1891–1907.
Clavel-Lévêque, M. (1989). La forêt gauloise vue des textes. Puzzle gaulois, les Gaules en
mémoire. Presses Universitaires de Franche-Comté, Besançon.
Closset-Kopp, D. and Decocq, G. (2015). Remnant artificial habitats as biodiversity islets into
forest oceans. Ecosystems, 18, 507–519.
Closset-Kopp, D., Hattab, T., Decocq, G. (2019). Do drivers of forestry vehicles also drive
herb layer changes (1970–2015) in a temperate forest with contrasting habitat and
management conditions? J. Ecol., 107(3), 1439–1456.
Dassié, J. (1978). Manuel d’archéologie aérienne. Technip, Paris.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
192 Historical Ecology

David, S. (2014). Rapport d’analyse et d’interprétation de données LiDAR ; forêts domaniales


de Compiègne et Laigue (Oise). Report, Office National des Forêts, Paris.
Delafosse, W. (1948). De l’origine des mardelles de Lorraine. Mémoire de l’Académie
nationale de Metz, 11, 63–85.
Etienne, D., Ruffaldi, P., Goepp, S., Ritz, F., Georges-Leroy, M., Pollier, B., Dambrine, E.
(2011). The origin of closed depressions in Northeastern France: A new assessment.
Geomorphology, 126, 121–131.
Feiss, T., Horen, H., Brasseur, B., Buridant, J., Gallet-Moron, E., Decocq, G. (2017a).
Historical ecology of lowland forests: Does pedoanthracology support historical and
archaeological data? Quat. Int., 457, 99–112.
Feiss, T., Horen, H., Brasseur, B., Lenoir, J., Buridant, J., Decocq, G. (2017b). Optimal
sampling design and minimal effort for soil charcoal analyses considering the soil type
and forest history. Veg. Hist. Archaeobot., 26, 627–637.
Frémont, J.M. and Woimant, B. (1975). Le vicus gallo-romain de la Carrière du Roi en forêt
de Compiègne (Oise), 1e partie. Revue archéologique de l’Oise, 6, 44–46.
Frémont, J.M. and Woimant, B. (1976). Le vicus gallo-romain de la Carrière du Roi en forêt
de Compiègne (Oise), 2e partie. Revue archéologique de l’Oise, 8, 45–49.
Georges-Leroy, M., Bock, J., Dambrine, E., Dupouez, J.L. (2011). Apport du lidar à la
connaissance de l’histoire de l’occupation du sol en forêt de Haye. Archéosciences, 35,
117–129.
Grégoire de Tours (1836). Historiae ecclesiasticae Francorum. In Histoire ecclésiastique des
Francs, livre IV, Guadet, J. (ed.). Société de l’histoire de France, Paris.
Leroyer, C., Barracand, G., Coussot, C., Werthe, E., Hulin, G., Jude, F., Leduc, C., Allenet de
Ribemont, G. (2017). Nouvelles données sur le paysage végétal de la moyenne vallée de
l’Oise au subboréal : l’étude de la séquence organique de Thourotte (Oise, France).
Quaternary, 28(4), 471–489.
Maury, A. (1848). Recherches historiques et géographiques sur les grandes forêts de la Gaule
et de l’ancienne France. In Mémoire de la Société des antiquaires de France, Duverger, E.
(ed). Paris.
Maury, A. (1850). Histoire des grandes forêts de la Gaule et de l’ancienne France. A.
Leleux, Paris.
Perrière, J. and Leroyer, C. (2006). Impact sur le milieu végétal et activités agropastorales des
groupes néolithiques d’après la séquence pollinique d’Armancourt (60). In Impacts
interculturels au Néolithique moyen. Du terroir au territoire : sociétés et espaces.
Duhamel, P. (ed). Revue Archéologique de l’Est, suppl. 25, Dijon.
Plue, J., Hermy, M., Verheyen, K., Thuillier, P., Saguez, R., Decocq, G. (2008). Persistent
changes in forest vegetation and seed bank 1,600 years after human occupation. Landsc.
Ecol., 23, 673–688.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
A Historical Ecology of the Compiègne Forest (N France) 193

Rackham, O. (1998). Savanna in Europe. In The Ecological History of European Forests,


Kirby, K.J. and Watkins, C. (eds). CABI Publishing, Wallingford.
Sittler, B. and Hauger, K. (2007). Les apports du laser aéroporté à la documentation de
parcellaires anciens fossilisés par la forêt : l’exemple des champs bombés de Rastatt en
Pays de Bade. In La mémoire des forêts, Dupouey, J.L., Dambrine, E., Dardignac, C.,
Georges-Leroy, M. (eds). ONF – INRA – DRAC Lorraine, Nancy.
Thuillier, P. (2017). Dynamique des paysages de l’Antiquité au Moyen Age. La naissance de
la forêt de Compiègne (Ier s. av. J.-C. – XIIIe s. ap. J.-C.). PhD Thesis, Université de
Picardie Jules Verne, Amiens.
Tuffreau-Libre, M. (1977). La céramique commune gallo-romaine de la forêt de Compiègne
(Oise) au Musée des Antiquités nationales. Cah. Archéol. Pic., 4, 125–160.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
15

The Chestnut Orchards in


the Bolognese Apennines: A
Vanishing Socio-ecological
Habitat
Giovanna PEZZI1, Fabrizio FERRETTI2, Alberto MALTONI3,
Patrik KREBS4, Marco CONEDERA4 and Giorgio MARESI5
1
University of Bologna, Italy
2
CREA Research Centre for Forestry and Wood, Arezzo, Italy
3
University of Florence, Italy
4
Swiss Federal Research Institute WSL, Cadenazzo, Switzerland
5
Fondazione Edmund Mach, San Michele all’Adige, Italy

15.1. Introduction

Castanea sativa Mill. (sweet chestnut) is a characteristic and iconic tree of the
southern European mountain landscape. Its presence has always been tightly linked
to human activities and the rural populations who cultivated the chestnut adopting
different management options (orchards, high forests and coppices). The first
evidence of a large-scale spread and cultivation of the species dates back to the
Romans, who were primarily interested in coppice management for wood (poles)
production (Conedera et al. 2004). Chestnut management for fruit production
increased in the post-Roman Era and in the Middle Ages, in particular when the
chestnut tree became the “bread tree” (albero del pane) in many mountain regions of
Europe, giving rise to what some people refer to as the “chestnut civilization”
(Gabrielli 1994). Existing chestnut groves still reflect the historical legacy of the

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
196 Historical Ecology

past importance of the tree: for instance, the high number of selected fruit varieties
(varying in ecological needs or ripening period), including nuts for flour production
as staple food or high-quality fruits (marroni and marroni-like) for seasonal
markets. Moreover, many phyto-toponyms related to this species and its cultivation
testify its presence throughout southern Europe (Squatriti 2013).

Figure 15.1. The study area and a view of an iconic traditional chestnut orchard.
For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip

First signs of decline of the Middle Age chestnut civilization occurred in the 18th
century as a result of the Little Ice Age climatic cooling, which caused widespread
frost damage to chestnut trees at the most exposed sites. In the following centuries,
other driving factors such as the improvement of agricultural cultivation techniques
and the introduction of new crops (e.g. maize and potatoes), the onset of new
diseases and the depopulation of the mountain countryside owing to industrialization
caused a progressive and generalized decline in traditional chestnut cultivation
(Conedera and Krebs 2008). In Italy, for instance, chestnut stands devoted to fruit
production (orchards) have decreased by about 90% since the beginning of the
20th century, passing from 608,000 ha (Vigiani 1908) to 60,000 ha (Bounous 2014).
Similarly, coppice stands underwent either an extension of the traditional rotation
time or total abandonment.

In this chapter, we focus on chestnut orchards of the Bolognese Apennines (Italy,


Figure 15.1), an iconic and ancient anthropogenic south-European landscape at risk
of vanishing (Squatriti 2013). In particular, we will apply a historical ecology
approach in order to analyze the past 80 years of its cultivation and propose possible
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Chestnut Orchards in the Bolognese Apennines 197

future management options in order to perpetuate this unique landscape feature also
in the future.

15.2. The traditional chestnut orchards

Traditional chestnut orchards (or groves) are open stands populated by grafted
C. sativa trees for fruit production. They are multipurpose systems (Mariotti et al.
2019) with a semi-natural herb layer usually devoted to pasture with domestic
animals (silvo-pastural systems, Figure 15.1). The presence of veteran trees
(sporadic in other forest formations) contributes to increase the conservation
importance of the habitat, as they preserve local varieties and provide favorable
living conditions (e.g. trunks with holes, dead wood, large crowns) for other
valuable taxa (animals, plants, lichens, mosses), including red list species (Krebs
et al. 2008; Pezzi et al. 2020a).

Figure 15.2. Conceptual framework of factors


affecting the chestnut orchard cultivation
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
198 Historical Ecology

The open structure and the simultaneous presence of woody elements and open
space herb species represent an ideal habitat for many plant and animal species (e.g.
Obrist et al. 2011; Morelli et al. 2019). As a result, traditional chestnut orchards are
biodiversity hotspots and habitats worthy of conservation (9,260 Castanea sativa
woods) according to the European Habitat Directive (92/43/EEC). The habitat value
is however the result of historical and socio-economic factors (e.g. ownership, food
needs) and externalities (e.g. conservation policies, tourism, fruit market) and may
thus vary highly in terms of the chestnut tree features (age, density, health, turnover,
pruning, variety composition), understory vegetation and related management
(mainly grazing and mowing), as well as ecological factors such as stand conditions
(climate, soil, topography, accessibility), pest and pathogen incidence and other
biotic interactions (e.g. game impact, propagule pressure of potential vegetation)
(Figure 15.2).

15.3. The chestnut groves of the Bolognese Apennines

The Bolognese Apennines are located to the southwest of the city of Bologna
(Lat. 44.492475; Long. 11.34297). They cover about 2,110 km2 and are
characterized by a marked altitudinal gradient from 49 to 1,944 m a.s.l. Chestnut
cultivation in this area has been rooted since the Middle Ages. On the basis of local
tradition, some authors even put forward the hypothesis that the chestnut-growing
tradition in this area was actively promoted by Countess Matilda of Canossa
(1046-1115), as demonstrated by the custom of designating old traditional chestnut
groves as Matildici (Alessandri et al. 2020). Chestnuts thus represented a vital staple
food for entire generations of mountain people and a commercial luxury product for
the nobility and the rich city dwellers of the area. As a result, the abundance of
chestnut orchards has been constantly linked to the demographic evolution, with
expansion during phases of economic growth and decrease in periods of depression,
such as the plague in the 14th century (Zagnoni 1997). The existence of both
chestnut (i.e. Pastonese cultivar) and marroni varieties is confirmed in historical
documents such as the chorographic dictionary by Calindri (1781–1783) and the
coeval Boncompagni cadaster (Pezzi et al. 2020b). At the beginning of the
20th century, chestnut-dominated formations covered ca. 12,863 ha of the area, at an
altitudinal range from 300 to 900 (1,000) m a.s.l., corresponding to the deciduous
Quercus-dominated and the Fagus sylvatica forest belts. Most of them (i.e.
10,928 ha) were classified as chestnut orchards, as reported in 1936 by the Italian
Kingdom Forest Map. This map is a very important historical reference allowing us
to estimate the dramatic decline of the chestnut orchards (at present covering 1,498 ha;
Regione Emilia-Romagna 2020) and their conversion into coppices or high forests
over the last 85 years (Ferretti et al. 2018).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Chestnut Orchards in the Bolognese Apennines 199

15.4. A changing world: abandonment, diseases and other problems

Similarly to other south European areas, the strategic role of chestnut culture was
subjected to a first serious threat with the appearance of the ink disease
(Phytophthora cambivora (Petri) Buism and Petri 1917) in the 19th century.
Existing concerns about this disease were confirmed by Quattrocchi (1938), who
found 1,024 ha of chestnut orchards affected by the ink disease in the Bologna
Apennines, together with 703 ha completely degraded and 7,862 ha in bad condition
because of the lack of appropriate management.

Following World War II, most Apennine villages witnessed a large migratory
movement towards cities and plains. The need to find a better way of life and to
escape from poverty and famine induced many young people to leave the mountains.
This wave of emigration, combined with the heavy war casualties, severely affected
the local communities in terms of both number of people and their links to the
territory and traditions (Gallingani 2019). The resulting social change produced a
progressive abandonment of chestnut culture and related activities such as tree
pruning, lawn mowing, collecting foliage and burs. Moreover, the contemporary
appearance and spread of the chestnut blight (Cryphonectria parasitica (Murr.)
Barr; Biraghi 1946) affected the trees’ health and fruit production, discouraging
most of the remaining chestnut growers: the “bread tree” was considered lost and the
chestnut era finished. Experts recommended coppicing the orchard stands (the
young sprouts were initially considered as more resistant to the disease; Biraghi
1953) or underplanting them with rapidly growing conifers. Paradoxically, the
possibility of selling the felled old and declining trees to tannin factories provided an
additional source of income and thus a further stimulus for leaving the mountain
regions. This abandonment mainly affected the upper Apennines, where most of the
chestnut orchards were devoted to flour production. On the contrary, the orchards
producing marroni continued to be cultivated, at least to some extent.

15.5. The turning point of the 1980s

In the 1980s, after decades of almost total abandonment of chestnut culture


culminating in the 1960s and 1970s, the situation began to change. The appearance
and natural spread of the chestnut blight hypovirulence, first observed in Italy at the
beginning of the 1950s (Biraghi 1953), allowed the chestnut orchard trees to survive
and even to recover from the blight attacks. This fungal infection, defined as a
healing canker, was able to colonize the bark and girdle the stem or the branch but
remained on the superficial layer of the bark, and the flow of the sap was not
affected, leaving the tree untouched and living. Over time, the healing canker
evolved and completely healed the original infection (Turchetti et al. 2008).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
200 Historical Ecology

Contrary to what happened with the American chestnut (Griffin 2000), the natural
diffusion of the hypovirulence and the healing form of the chestnut blight gave rise
to a new and very optimistic scenario for the European chestnut and its cultivation,
particularly in Italy.

Moreover, some emigrants returned home after retirement and began restoring
the chestnut orchards and introducing new practical and technological management
and marketing strategies, including the birth of growers’ associations to sustain the
sector. The restoring approach consisted of removing the invading vegetation and
pruning the old and grafted chestnut trees. New plantations and, more especially,
conversions of coppices into orchards by means of grafting the resprouts (Gherardi
et al. 1991) also took place, resulting in a significant increase in chestnut cultivation.
In this context, the commercial and valuable marroni played a central driving role
thanks to their appreciation in urban markets as a high-quality natural product and
the related optimal price/performance ratio. Furthermore, when selling chestnut
fruits, people also market the traditional culture and the related bucolic landscape,
which is indeed far removed from the original and grim reality of the chestnut as the
tree of bread. As a result, the restoring activities mostly involved the marroni
orchards in the lowlands of the Apennines. Interestingly, this happened much less in
the upper part of the mountains, where the traditional staple food orchards are
located.

15.6. Current constraints and future perspectives

The chestnut tree is currently facing old and new threats (Figure 15.3).
Fortunately, the hypovirulence of the chestnut blight is still predominant throughout
the whole chestnut range, permitting the survival of trees and their cultivation (Pezzi
et al. 2011, 2020a). The ink disease is still widespread in the originally affected
areas (unpublished data), but a further spread in terms of new foci is very limited
and localized (Ambrosini et al. 1997; Pezzi et al. 2011). This suggests that the
pathogen does not have the potential to spread into the whole chestnut area as
previously hypothesized. A new threat to the chestnut tree appeared in Europe at the
beginning of the 21st century (Brusino et al. 2002) and spread to the Bolognese
Apennines starting from 2008 (Graziosi and Santi, 2008): the Asian chestnut gall
wasp (Dryocosmus kuriphilus Yasumatsu). Even if no trees were directly killed by
this pest, the heavy damages to the foliage caused a drastic reduction in fruit
production, again turning the confidence of the growers into very pessimistic
scenarios. The biological control options with the specific parasitoid Torymus
sinensis Kamijo (Quacchia et al. 2008) have fortunately proved to be very effective
since the biological agent was first released in 2010 (Vai et al. 2014). Damage
caused by the chestnut gall wasp is at present very limited and does not affect the
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Chestnut Orchards in the Bolognese Apennines 201

trees’ vitality or the fruit production (Pezzi et al. 2020a). Therefore, the management
of the main chestnut phytosanitary issues can currently be carried out in a
completely biological and sustainable way.

Figure 15.3. Summary of the historical evolution of chestnut orchards


and its relationship with the different constraints in the Bolognese
Apennines (modified from Pezzi et al. 2011)

A further new risk factor for the chestnut is climate change in general and the
ensuing climatic extremes such as summer drought in particular (Conedera et al.
2010). Drought affecting the fruit production occurred in this part of the Apennines
in 2017 and 2019. In addition, the chestnut groves were affected by a snowstorm in
2017. Such repeated climatic stress can reduce the resistance of the trees against
diseases or cause localized decline phenomena. In this context, the nut rot fungus
(Gnomoniopsis castaneae Tamietti) newly appeared, severely affecting the fruit
production (especially in 2018). Its spread seems to be strictly related to the
temperature increase during the vegetative season and particularly at the time of the
fruit harvest.

From a socio-economic point of view, the restoration and maintenance of


chestnut orchards and their ecological value is closely linked to the presence of
interested owners carrying out traditional management. At present, 115 growers
mainly located in the lower parts of the mountains and dedicated to marroni
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
202 Historical Ecology

valorization are reported as affiliated to the local chestnut growers’ association.


Interestingly, a generational change is taking place, with an increasing number of
young members. In the upper part of the range, there is a small consortium of about
30 members dedicated to chestnut cultivation and flour production. The number of
consortia members has changed over the years due to the possibility of easy access
to financial support from public bodies. However, there is also the effect of the
aging of the growers, with an old generation not always being replaced and the
subsequent abandonment of the orchards (Pezzi et al. 2017). In addition, the upper
Apennines are still affected by a demographical decrease. On the other hand, the
recent rediscovery of the value of flour and the appearance of new products related
to chestnut, such as aromatized beers, are opening interesting new scenarios also for
the economy of the old flour orchards. Nevertheless, the firms related to agriculture
or silviculture in the Apennines show a constant trend of reduction, although slightly
slowing down in recent years.

In summary, chestnut cultivation is still vital, but at risk of marginalization and


subjected to a balancing act between abandonment and making the recognized
multifunctionality potential for the mountains effective. Although the traditional
management of chestnut orchards allows a high level of biodiversity to be preserved,
there is still a substantial risk of abandonment of these activities, which is related to
the low marketing possibilities of products, the highly fragmented property systems,
and the aging of the growers and difficulties in granting a management succession.
The recognition of the indispensable contribution afforded by chestnut growers
should lead to the definition of political choices for territorial governance aimed at
supporting, also financially, the work of these subjects. Currently, there is still an
important and significant economic income in some areas, but often carried out
through personal passion or transmitted between generations as a family cultural
heritage. Conserving the multifunctional values of chestnut groves in the Bolognese
Apennines thus calls for a complex management strategy that comprises:
(i) supporting young growers and helping them to gain possession of chestnut
orchards; (ii) supporting associations that enhance the political weight of growers
and contribute to spreading traditional and innovative knowledge on chestnut
orchards management; (iii) increasing research on both the physiology and ecology
of the chestnut and the orchard trees in particular, in order to better cope with future
challenges such as climate change and evolving socio-economic conditions, and
(iv) exploring new strategies to give added value to the chestnut fruit products in the
context of a sustainable management.

A specific effort to safeguard old-growth chestnut orchards could provide a


representative example and encourage increased attention to chestnut cultivation,
assuring the revitalization of this fundamental heritage of the Italian mountains.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Chestnut Orchards in the Bolognese Apennines 203

15.7. References

Alessandri, S., Krznar, M., Ajolfi, D., Cabrer, A.M.R., Pereira-Lorenzo, S., Dondini, L.
(2020). Genetic diversity of Castanea sativa Mill. accessions from the Tuscan-Emilian
Apennines and Emilia Romagna Region (Italy). Agronomy, 10, 1–11.
Ambrosini, I., Gherardi, L., Viti, M.L., Maresi, G., Turchetti, T. (1997). Monitoring diseases
of chestnut stands by small format aerial photography. Geocarto International, 12(3),
41–46.
Biraghi A. (1946). Il cancro del castagno causato da Endothia parasitica. L’Italia Agric., 7,
406.
Biraghi, A. (1953). Possible active resistance to Endothia parasitica in Castanea sativa.
Proceedings. of the XI Congress of the International Union of Forest Research
Organizations, 643–645.
Bounous, G. (2014). Il castagno. Risorsa multifunzionale in Italia e nel mondo. Edagricole,
Bologna.
Brussino, G., Bosio, G., Baudini, M., Giordano, R., Ramello, F., Melika, G. (2002).
Pericoloso insetto esotico per il castagno europeo. L’Informatore Agrario, 37, 59–61.
Conedera, M. and Krebs, P. (2008). History, present situation and perspective of chestnut
cultivation in Europe. In Proceedings of the Second Iberian Congress on Chestnut, Abreu,
C.G, Peixoto, F.P., Gomes-Laranjo J. (eds). ISHS, Section Nuts and Mediterranean
Climate Fruits, Leuven.
Conedera, M., Krebs, P., Tinner, W., Pradella, M., Torriani, D. (2004). The cultivation of
Castanea sativa (Mill.) in Europe, from its origin to its diffusion on a continental scale.
Veg. Hist. Archaeobot., 13, 161–179.
Conedera, M., Barthold, F., Torriani, D., Pezzatti, G.B. (2010). Drought sensitivity of
Castanea sativa: Case study of summer 2003 in the Southern Alps. Acta Hortic., 866,
297–302.
Ferretti, F., Sboarina, C., Tattoni, C., Vitti, A., Zatelli, P., Geri, F., Pompei, E., Ciolli, M.
(2018). The 1936 Italian kingdom forest map reviewed: A dataset for landscape and
ecological research. Ann. Silvic. Res., 42(1), 3–19.
Gabrielli, A. (1994). La civiltà del castagno. Monti e boschi, 65, 3.
Gallingani M.A. (2019). Rapporto Appennino 2019 [Online]. Available at: http://inumeridibo
lognametropolitana.it/studi-e-ricerche/rapporto-appennino-2019) [Accessed 27 January 2021]
Gherardi, L., Maetzke, F., Maresi, G. (1991). Prove sperimentali di recupero produttivo dei
castagneti da frutto nella collina bolognese. L’Italia Forestale e Montana, 46(2),
159–173.
Graziosi, I. and Santi, F. (2008). Chestnut gall wasp (Dryocosmus kuriphilus): Spreading in
Italy and new records in Bologna province. Bull. Insectol., 61(2), 343–348.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
204 Historical Ecology

Griffin, G.J. (2000). Blight control and restoration of the American chestnut. J. Forest.,
98(2), 22–27.
Krebs, P., Moretti, M., Conedera, M. (2008). Castagni monumentali nella Svizzera sudalpina.
Importanza geostorica, valore ecologico e condizioni sanitarie. Sherwood, 14(1), 5–10.
Mariotti, B., Castellotti, T., Conedera, M., Corona, P., Manetti, M.C., Romano, R., Tani, A.,
Maltoni, A. (2019). Linee guida per la gestione selvicolturale dei castagneti da frutto.
Rete Rurale Nazionale 2014–2020, Scheda 22.2 – Foreste, CREA, Rome.
Morelli, F., Python, A., Pezzatti, G.B., Moretti, M. (2019). Bird response to woody pastoral
management of ancient chestnut orchards: A case study from the southern Alps. For.
Ecol. Manage., 453, 117560.
Obrist, M.K., Rathey, E., Bontadina, F., Martinoli, A., Conedera, M., Christe, P., Moretti, M.
(2011). Response of bat species to sylvo-pastoral abandonment. Forest Ecol. Manag.,
261(3), 789–798.
Petri, L. (1917). Studi sulla malattia del castagno detta “dell’inchiostro”. Ricci, M.,
Florence.
Pezzi, G., Maresi, G., Conedera, M., Ferrari, C. (2011). Woody species composition of
chestnut stands in the Northern Apennines: The result of 200 years of changes in land use.
Landsc. Ecol., 26(10), 1463–1476.
Pezzi, G., Lucchi, E., Maresi, G., Ferretti, F., Viaggi, D., Frascaroli, F. (2017). Abandonment
or survival? Understanding the future of Castanea sativa stands in function of local
attitude (Northern Apennine, Italy). Land Use Policy, 61, 564–574.
Pezzi, G., Gambini, S., Buldrini, F., Ferretti, F., Muzzi, E., Maresi, G., Nascimbene, J.
(2020a). Contrasting patterns of tree features, lichen, and plant diversity in managed and
abandoned old-growth chestnut orchards of the northern Apennines (Italy). Forest Ecol.
Manage., 470–471, 118207.
Pezzi, G., Donati, D., Muzzi, E., Conedera, M., Krebs, P. (2020b). Using chorographic
sources to reconstruct past agro-forestry systems. A methodological approach based on
the study case of the northern Apennines. Landscape Res., 45, 359–376.
Quacchia, A., Moriya S., Bosio, G., Scapin, I., Alma, A. (2008). Rearing, release and
settlement prospect in Italy of Torymus sinensis, the biological control agent of the
chestnut gall wasp Dryocosmus kuriphilus. Biol. Control, 53, 829–839.
Quattrocchi, G. (1938). Il miglioramento dei castagneti dell’Appennino Bolognese.
Stabilimento grafico F. Lega, Faenza.
Squatriti, P. (2013). Landscape and Change in Early Medieval Italy: Chestnuts, Economy,
and Culture. Cambridge University Press, Cambridge and New York.
Turchetti, T., Ferretti, F., Maresi, G. (2008). Natural spread of Cryphonectria parasitica and
persistence of hypovirulence in three Italian coppiced chestnut stands. Forest Pathol., 38,
227–243.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The Chestnut Orchards in the Bolognese Apennines 205

Vai, N., Colla, R., Mazzoli, L., Bariselli, M. (2014). The regional project for biological
control of the Chinese gall wasp Dryocosmus kuriphilus in Emilia-Romagna. In
Conference and Abstracts Book of the European Conference of Arboriculture − Planning
the Green City: Relationships Between Trees and Infrastructures, Giordano, L., Ferrini,
F., Gonthier, P. (eds). DISAFA Editions, Grugliasco.
Vigiani, D. (1908). Il castagno. Casa Agricola Fratelli Ottavi, Casale Monferrato.
Zagnoni R. (1997). La coltivazione del castagno nella montagna fra Bologna e Pistoia nei
secoli XI–XIII. Atti delle Giornate di Studio (Capugnano, 14 settembre 1996), Gruppo di
studi alta valle del Reno-Società pistoiese di storia patria (“Storia e ricerca sul campo fra
Emilia e Toscana”, 5), 41–57.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
16

Claudius’ Coin in the


Forest – Niche Construction
and Strategies by Early
Colonizers of Boreal Inlands
in Central Scandinavia
Ove ERIKSSON1 and Karl-Johan LINDHOLM2
1
Stockholm University, Sweden
2
Uppsala University, Sweden

16.1. Introduction

At the time when the first emperors ruled the Roman Empire around 2,000 years
ago, a colonization process commenced in the vast boreal forest inland regions of
central Scandinavia (Figure 16.1). The people involved in this colonization may
have lived in the remote outskirts of the inhabited world, but they had knowledge of
agriculture and iron production technology, and they most likely held a world-view
not so different from other people closer to the Roman frontier. Perhaps they had
heard something about a mighty empire far away in the south; at least someone had,
as a coin from the time of Emperor Claudius (AD 41–54) has been retrieved from a
lump of earth found in a slope towards the river Indalsälven near the small village of
Krångede (Figure 16.1). In this chapter, we examine this colonization, focusing on
the challenges faced by people, the strategies used to meet these challenges, and
how the colonization altered the landscape and vegetation. Due to the still limited
evidence, our examination should be regarded as tentative and our conclusion as a

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
208 Historical Ecology

hypothesis. The time frame is from early Iron Age (the Scandinavian Iron Age is
usually defined as the period 500 BC–AD 1050) until the early Middle Ages, but we
also briefly remark on later history and legacies in the present-day landscape in the
region.

Figure 16.1. A map showing Scandinavia, with the “boreal forest inlands regions of
central Scandinavia” hatched. The circles mark the closest main agricultural regions
during the early Iron Age. The dotted line marks the southern border of present-day
boreal zone. The dot marks Krångede, the place where Claudius’ coin was found.
For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Claudius’ Coin in the Forest – Niche Construction 209

In southern and south-eastern Scandinavia about 2,000 years ago, there were
extensive cultural landscapes with roots in the Neolithic, several thousand years
earlier. Forests had been cleared, and particularly on the plains, landscapes were
more or less fully exploited, with settlements, fields, meadows and pastures. Iron
tools were commonly available. At this time, land-use was organized according to
so-called “infield systems”. The infields – i.e. settlements, crop fields and meadows
– were protected (by fences or stone walls) from uncontrolled grazing by livestock,
and surrounded by “outland”, mainly used for livestock grazing and collection of
various resources (Pedersen and Widgren 2011; Eriksson and Arnell 2017; Eriksson
et al. 2021). During the first centuries AD, several “power centers” developed,
probably local “elite” chiefdoms, from which the colonization processes emanated,
for example south-eastern Sweden, the Gulf of Bothnia and western Norway (Figure
16.1).

Figure 16.2. A typical view of the boreal forest inland region of central Scandinavia,
showing a shieling (secondary farm) in the province of Jämtland. Photo by Andreas
Hennius. For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip

In conventional historical narratives, the boreal forest inland regions of central


Scandinavia have been considered as being more or less unaffected by people until
the Middle Ages (e.g. Myrdal 2011), only inhabited by sparse populations of
hunters. This view has been questioned, and there is evidence that the regions were
colonized much earlier (e.g. Magnusson 1986; Lindholm et al. 2013) and became
integrated parts of trade networks (e.g. von Stedingk and Baudou 2006), primarily
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
210 Historical Ecology

with iron (Hyenstrand 1979; Magnusson 1986; af Geijerstam and Nisser 2011), and
also with products from hunting (Lindholm and Ljungkvist 2016; Hennius 2020)
and tar (Hennius 2018). Similarly, in biology, a common narrative is that boreal
inland forests were largely “untouched” until what we may call modern forestry
began during the 19th century (see Eriksson (2018), for a review). Several in-depth
studies of forest history have shown that this narrative is flawed (e.g. Josefsson et al.
2010). However, the people colonizing these boreal forest inland regions during the
Iron Age and the Middle Ages would have faced great challenges. The climate is
harsh and except for some areas, for example along rivers and plains surrounding
large lakes, most land consists of coniferous forest and mires not primarily suitable
for agriculture (Figure 16.2). The main objective of this chapter is to examine the
strategies people used to handle these challenges, and how these affected the
landscape.

16.2. Concepts and theoretical framework

In our studies of the colonization of the boreal forest inland, we use a conceptual
and theoretical framework, human niche construction, useful for examining
continuous interactions between humans and their environment. The underlying idea
is that human actions, intentional or unintentional, alter the environment in a way
that feedback to people, either enabling them to increase their resource base,
promote inventions or constraining their further actions. Over time, these continuous
interactions may lead to cultural “adaptations”, in this context meaning perceptions
and practices which promote the persistence of local societies (for an extensive
discussion of these concepts, see Eriksson et al. 2018). Niche construction can be
defined broadly as the processes by which a species (e.g. humans) alters its own
ecological niche, or the niche of other species, and the feedback of these alterations
to the niche constructing species (Odling-Smee et al. 2003). The “ecological niche”
here means the totality of the resources and environmental requirements used by the
niche constructing species. “Niche” therefore equals “the environment”, or some
part of it. A key point in niche construction is that the species not only respond to
the environment (as it is often considered), but also in fact creates its own
environment, for good, or for bad.

Niche construction theory was originally developed in biology (Odling-Smee


et al. 2003), but has increasingly been applied to interactions between humans and
environment (e.g. Laland and O’Brien 2012; Eriksson et al. 2021), for example
concerning domestication of plants and animals (e.g. Zeder 2016). Indeed, on a large
spatial scale, we may consider human-mediated transformation of the environment
as a domestication of whole landscapes. Over time there is an “ecological memory”
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Claudius’ Coin in the Forest – Niche Construction 211

of past environmental alterations, and this ecological memory is what studies of


current landscapes and vegetation recognize as legacies of previous land-use.

16.3. A historical overview of the colonization

It is likely that the initial driver behind the colonization of the boreal forest
inland regions during the first centuries AD was iron production. Iron was produced
in bloomery furnaces using lake or bog ore (af Geijerstam and Nisser 2011). In just
one of the present-day provinces of the region (Jämtland-Härjedalen), over 700 iron
production sites have been found, and it is likely that the true number is higher
(Magnusson 1986). Based on remains of slag, Magnusson (1986) found that the
earliest iron production sites in the boreal forest inland regions (up until around the
mid first millennium AD) were considerably larger than later production sites (from
AD 900 onwards). This suggests that there was an “overproduction” of iron used for
trade during the first centuries AD, in line with what would be expected if the
colonization was initiated, directly or indirectly, by the Iron Age “chiefdoms”
located west, east and southeast of the inland forest regions (Figure 16.1). After a
period of decline around the 7th–8th century AD, the colonization of the boreal
forest inland regions increased during late Iron (Viking) Age and early Middle Ages
(e.g. Myrdal 2011), probably at least partly driven by the same mechanism, iron
production used for trade. However, the development of the blast furnace
technology during the early Middle Ages made other regions more efficient in
producing iron, and the smaller iron production sites in the inland forest regions at
that time indicate that they may have served only local demands. The colonization
was halted during the “late medieval agrarian crisis”, which followed after a series
of plague epidemics (mainly the “Black Death” in mid-14th century), wars and
climate cooling. Many farms were abandoned. Along with the recovery from these
societal crises, the inland forests in central Sweden may have been particularly
attractive to colonizers, since previously abandoned farms in more central regions
had been used for pasture or meadow, and thus became unavailable to people when
the population started to increase (Myrdal 2011). During the 18th and 19th centuries,
the number of small farms increased as a result of a growing population and
initiatives by the Swedish State to promote agricultural expansion (Gadd 2011).
Many of these small farms were later abandoned, as rural people moved into urban
centers for work in industries.

After this brief overview, we return to the earlier phases of the colonization
process, initiated during the Iron Age. Before focusing on the strategies employed
by people, it is relevant to consider what kind of knowledge, and what kind of
world-view, “cosmology”, the colonizers may have brought with them. It is likely
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
212 Historical Ecology

that the people colonizing the boreal forest inland regions of Scandinavia had similar
knowledge and world-view as people from the main agricultural regions (Lindholm
et al. 2013). It seems evident that people must have known iron production
technology as the earliest iron production sites were located further to the south in
comparison with the forest inland regions (af Geijerstam and Nisser 2011).
Furthermore, infield systems developed the centuries before and after AD (Eriksson
2020). These systems included the conception of land as infields (fields, meadows),
probably “owned” by persons or families, and outland, which was used as commons,
shared by several farms. Furthermore, livestock (cattle, sheep and goats) were the
basis of the agricultural system, since manure was needed to fertilize the small
permanent crop fields. Livestock was commonly kept indoors (“stalled”), parts of
the year. Stalling may have been introduced for various reasons, for example
protection from theft and to enhance collection of manure, but may eventually have
served to protect livestock from harsh winters (Eriksson 2020). Thus, fodder
production on meadows (hay or leaf-hay) was an essential component of the
agricultural system. From more recent times, we know the yearly cycle of
management of wooded meadows adapted to maximize the production of hay
(including leaf-hay), but it is likely that these management practices were established
during the early Iron Age, along with the invention of scythes and hay-rakes useful
for hay harvest (Eriksson 2020). The links from plant biomass (fodder), not
accessible as food for humans, via manure, to fertilization of crop fields, were
understood (although “understood” would mean something very different for people
2,000 years ago, who had no knowledge about chemistry or nutrients).

16.4. A structured landscape

16.4.1. Constructing the environment

Several components of infield systems were literally the result of niche


construction, for example crop production on manured permanent fields and meadow
management, and these formed a basis of conceptions of how land was managed and
organized, and probably also about who had the rights to use the land, an early form of
“private ownership” (Eriksson and Arnell 2017). Although hunting, for example
Eurasian elk and reindeer (Hennius 2020), and fishing were probably part of the basis
for subsistence during the colonization, people came from agricultural regions, i.e.
they were farmers. We should imagine that people involved in the colonization of the
inland forests had a mind-set based on infield systems with their constructed
production sites, fields and meadows, and where stalling of livestock was part of the
yearly cycle, and where herded livestock during the vegetation period grazed in
semi-open forests (as they become when livestock grazes forests).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Claudius’ Coin in the Forest – Niche Construction 213

These people now faced a landscape with large tracts of boreal forest and mires
and where only minor areas along rivers and large lakes were initially suitable for
agriculture. A constraint was also that iron production sites had to be chosen where
iron ore is available, and these sites were not necessarily most suitable for
settlements (Magnusson 1986). Thus, the use of the landscape had to be spatially
structured into areas for subsistence based on agriculture and areas for iron
production. We suggest that the structuring of the land-use, which initially was a
“problem” to solve, represents an overarching strategy employed by people to
sustain their livelihood under these conditions. The landscape as a whole became
domesticated, and different components of this domestication were result of niche
construction. We stress that when we use the term “strategy”, we do not imply that
this was planned with foresight, that anyone cleverly thought it out from the
beginning. Strategies develop over time, as a result of a combination of an initial
mind-set, necessities due to constraints, accumulated experience and knowledge
transfer, and they may have been partly unintentional byproducts of decisions
initially made for other purposes.

16.4.2. Managing livestock

Forest grazing by livestock has been used as long as agriculture has existed in
Scandinavia, and it was a component of infield systems as well, i.e. grazing was
conducted in the outland, mostly forests. Unmanaged boreal forests are not very
suitable for livestock grazing, as the major tree species, Scots pine (Pinus sylvestris)
and Norway spruce (Picea abies), and the dominating field layer composed of
ericaceous shrubs are not sufficient feed for cattle and sheep; goats may have thrived
better, and it is possible that goats were mostly used initially. However, if the forest is
cleared, for example for collection of wood for fuel or charcoal, and especially if it is
burned, more attractive feed is promoted, such as the grass Deschampsia flexuosa,
herbs such as Chamaenerion angustifolium, and deciduous species such as birch
(Betula spp.), goat willow (Salix caprea) and rowan (Sorbus aucuparia) (e.g. Groven
and Niklasson 2005). Burning to clear forests and improve forest grazing is likely to
have been used since long in southern Sweden. This management may therefore have
been part of the colonizer’s initial knowledge, although it was here placed in a
different environmental context. The natural fire frequency of boreal forests would
thus have increased due to anthropogenic impact (e.g. Granström and Niklasson 2008).

For southern Sweden during the 17th century onwards, Dahlström (2006)
estimated that livestock density in forests could be as low as c. 0.1 cattle/hectare.
The figure is likely to be even lower in northern boreal forest because these are less
productive. Considering that cattle which graze in forests move up to eight
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
214 Historical Ecology

kilometers per day, large areas had to be burnt in order to sustain even quite small
herds of livestock. For climatic reasons in the forest inland regions, livestock needed
to be hosted indoors during the coldest and wettest parts of the year, probably at
least seven to eight months. Large quantities of fodder were needed. Initially, and
for large parts of Scandinavia, leaf-hay was used as fodder, i.e. leaves and twigs
from deciduous tree species (Slotte 2001). As the dominating tree species in boreal
forest, Scots pine and Norway spruce, are poor as fodder, also fodder production
required active manipulation of the forest environment. Again, burning would be the
best way to achieve a decent production of fodder, both hay (grass, herbs) and
leaf-hay. However, this required that the area used for fodder production could not
simultaneously be used for grazing. As stalling of livestock was done at the farm,
and in order to reduce the need of hay-transport, it is reasonable that the area closest
to the settlement was allocated to fodder production. This means that the constructed
landscape resembled the structure of infield systems, although the spatial scale was
considerably larger, as the overall productivity of the boreal forest is so much
smaller than on the agricultural plains. Livestock should then be herded at a
distance, perhaps even a considerable distance, from the settlements.

In the boreal forest landscape, mires (including bogs) cover vast areas, and these
provide an additional source of fodder, mainly sedges (Carex spp.). Management of
outland hay-meadows on mires is well known from more recent times, when
management often included manipulation of the water-regime in order to promote
hay production (e.g. Elveland 2015). Harvesting hay on wetlands is extremely
labor-intensive, and the manpower available for such harvest is likely to have been a
constraint. Furthermore, as with iron production sites, the sites with the most useful
mires may not be close to the settlements, implying a need for long transports of the
hay after harvest. Thus, we recognize that people may have experienced several
constraints. Herding of livestock (at distances away from settlements) and
hay-harvest were time-consuming. In addition, hay harvested far from settlements
must either be transported directly after harvest or stored until transport during
winter. These activities necessitated an extension of the spatial scale of the area
needed for the capturing of resources needed for subsistence; here, we also remind
that iron production sites were located away from settlements. We suggest that these
constraints led to the development of an overarching strategy for subsistence in the
boreal forest inland: shielings.

16.4.3. Shielings (secondary farms)

Shielings are secondary farms, belonging to a main farm, but located at some
distance away and used during parts of the year, i.e. a kind of transhumance system
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Claudius’ Coin in the Forest – Niche Construction 215

where some people and most livestock moved away from the main farm in order to
extend the resource base for subsistence (e.g. Svensson 2018). Shielings are usually
placed in the context of farming systems from more recent times, and based on a
19th century model of shielings, Larsson (2009) argued that these originated no
earlier than the late Middle Ages. Other authors suggest that shielings originated
during the Iron Age (e.g. von Stedingk and Baudou 2006). Based on the constraints
described above, we agree with the latter opinion, and suggest that the places that
became shielings developed as a cultural adaptation to the specific conditions facing
people during colonization of the forest inland.

Equipped with a mind-set based on infield systems, people may have strived to
construct the landscape in that way. However, the poor production of boreal forests
necessitated clearing and burning of large areas. The area closest to settlements had
to be allocated for production of leaf-hay (or hay from grass and sedges if
settlements were located close to rivers). In addition, extensive areas were needed
for outland grazing. It is thus easy to imagine that an idea developed to organize
summer grazing at some distance from the main farm, thus relieving the herders
from moving livestock daily between the main farm and the grazing grounds, and
ensuring that livestock did not destroy the production of hay in the vicinity of
settlements. A general explanation for the origin of shielings would be that they
were adaptive in a resource-poor landscape, as a means to increase the exploitation
of fodder by seasonally moving livestock beyond the areas close to the infields. This
would seem as a natural thing to do, if we assume that the colonizers brought with
them the knowledge of infield systems. Karlsson et al. (2010) suggested that the
shieling-farm connection also served to define an area that demonstrated the extent
of the farm’s territory, thus preventing other colonizers from establishing there.

We still lack of clear material evidence for early shielings, but as pointed out by
Lindholm et al. (2013), remains of early shielings may be archaeologically obscure.
Magnusson (1986) made the interesting remark that iron production was often
located at sites which seemed as suitable sites for shielings. We may speculate that
the clearing of forest for producing charcoal used in iron production created
preconditions for the localization of an early form of shielings; the forest areas
cleared for charcoal production were used for livestock grazing. This would also
have “rationalized” the guarding of the livestock herd, since people were engaged in
other activities at the same site. In fact, Emanuelsson and Segerström (2002)
suggested that a mutual relationship ultimately developed between charcoal
production, slash-and-burn agriculture (which probably was introduced during late
Middle Ages) and livestock grazing.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
216 Historical Ecology

16.5. Concluding remarks

Figure 16.3. A shieling from the province of Värmland. This shieling is


still managed, as a biological cultural heritage. For a color version
of this figure, see www.iste.co.uk/decocq/ecology.zip

The colonization and establishment of agriculture and iron production in the boreal
forest inland regions of central Scandinavia from the Iron Age onwards implied that
people must have overcome great difficulties. Climate is harsh, and vast areas are
covered with forest and mires initially unsuitable for farming. We suggest that a
combination of a mind-set based on infield systems occurring in the regions from
where the colonizers came, and constraints imposed by the environment, promoted
human niche construction processes specific for these regions. The landscape was
basically structured as infields and outland, but the constraints imposed by the poor
productivity of the forests necessitated extending forest grazing spatially, promoted
use of outland mires for harvest of winter fodder, and ultimately led to development of
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Claudius’ Coin in the Forest – Niche Construction 217

secondary farms, shielings, as an innovative strategy of a spatially structured


domesticated landscape. Although initially developed during the Iron Age, other
processes transformed shielings to what they became later during history. Overlaid
with more recent use of boreal forests, mainly modern forestry, there are still
remaining legacies in the landscape, for example remaining shielings (Figure 16.3),
and rare cases of a forest structure that mirrors former use of forests as grazing ground,
with large Scots pines, grass and herbaceous plants (e.g. Ericsson et al. 2000). The
concept of the boreal forest regions as a cultural landscape has not been much
acknowledged in research, and we end with a hope that this landscape’s more than
2,000-year long history of human impacts should receive the attention it deserves.

16.6. References

Dahlström, A. (2006). Pastures, livestock number and grazing pressure 1620–1850.


Ecological aspects of grazing history in south-central Sweden. Thesis, Acta Universitatis
Agriculturaea Suecica 2006:95 [In Swedish with English summary].
Elveland, J. (2015). Wet hay-meadows in northern Sweden – Their history and present status.
Svensk Bot. Tidskr., 109, 292–336 [In Swedish with English summary].
Emanuelsson, M. and Segerström, U. (2002). Medieval slash-and-burn cultivation: Strategic
or adapted land use in the Swedish mining district. Environ. Hist., 8, 173–196.
Ericsson, S., Östlund, L., Axelsson, A.-L. (2000). A forest of grazing and logging:
Deforestation and reforestation history of a boreal landscape in central Sweden. New
Forests, 19, 227–240.
Eriksson, O. (2018). What is biological cultural heritage and why should we care about it? An
example from Swedish rural landscapes and forests. Nat. Conserv., 28, 1–32.
Eriksson, O. (2020). Origin and development of managed meadows in Sweden: A review.
Rural Landsc. – Soc., Environ., Hist., 7(2), 1–23.
Eriksson, O. and Arnell, M. (2017). Niche construction, entanglement and landscape
domestication in Scandinavian infield systems. Landsc. Res., 42, 78–88.
Eriksson, O., Ekblom, A., Lane, P., Lennartsson, T., Lindholm, K.-J. (2018). Concepts for
integrated research in historical ecology. In Issues and Concepts in Historical Ecology:
The Past and Future of Landscapes and Regions, Crumley, C.L., Lennartsson, T., Westin,
A. (eds). Cambridge University Press, Cambridge.
Eriksson, O., Arnell, M., Lindholm, K.-J. (2021). Historical ecology of Scandinavian infield
systems. Sustainability, 13, 817.
Gadd, C.-J. (2011). The agricultural revolution in Sweden 1700–1870. In The Agrarian
History of Sweden: 4000 BC to AD 2000, Myrdal, J. and Morell, M. (eds). Nordic
Academic Press, Lund.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
218 Historical Ecology

af Geijerstam, J. and Nisser, M. (eds) (2011). Swedish Mining and Metalworking. National
Atlas of Sweden. Norstedts Förlagsgrupp, Stockholm.
Granström, A. and Niklasson, M. (2008). Potentials and limitations for human control over
historic fire regimes in the boreal forest. Phil. Trans. R. Soc. B, 363, 2353–2358.
Groven, R. and Niklasson, M. (2005). Anthropogenic impact of past and present fire regimes
in a boreal forest landscape of southeastern Norway. Can. J. Forest Res., 35, 2719–2726.
Hennius, A. (2018). Viking Age tar production and outland exploitation. Antiquity, 92,
1349–1361.
Hennius, A. (2020). Towards a refined chronology of prehistoric pitfall hunting in Sweden.
Eur. J. Archaeol., 23, 530–546.
Hyenstrand, Å. (1979). Iron and iron economy in Sweden. In Iron and Man in Prehistoric
Sweden, Clarke, H. (ed.). Jernkontoret, LTs Förlag, Stockholm.
Josefsson, T., Gunnarsson, B., Liedgren, L., Bergman, I., Östlund, L. (2010). Historical
human influence on forest composition and structure in boreal Fennoscandia. Can. J.
Forest Res., 40, 872–884.
Karlsson, H., Emanuelsson, M., Segerström, U. (2010). The history of a farm-shieling system
in the central Swedish forest region. Veg. Hist. Archaeobot., 19, 103–119.
Laland, K.N. and O’Brien, M.J. (2012). Cultural niche construction: An introduction. Biol.
Theory, 6, 191–202.
Larsson, J. (2009). Summer farms in Sweden 1550 to 1920: An important element in North
Sweden’s agricultural system. Thesis, Jamtli Förlag, Östersund [In Swedish with English
summary].
Lindholm, K.-J. and Ljungkvist, J. (2016). The bear in the grave: Exploitation of top predator
and herbivore resources in first millennium Sweden – First trends from a long-term
research project. Eur. J. Archaeol., 19, 3–27.
Lindholm, K.-J., Sandström, E., Ekman, A.-K. (2013). The archaeology of the commons.
J. Archaeol. Anc. Hist., 10, 1–49.
Magnusson, G. (1986). Prehistoric and Medieval Iron Industry in the Province of Jämtland.
Jernkontorets Bergshistoriska Skriftserie N:r 22. Jernkontoret, Stockholm [In Swedish
with English summary].
Myrdal, J. (2011). Farming and feudalism 1000–1700. In The Agrarian History of Sweden:
4000 BC to AD 2000, Myrdal, J. and Morell, M. (eds). Nordic Academic Press, Lund.
Odling-Smee, F.J., Laland, K.N., Feldman, M.W. (2003). Niche Construction: The Neglected
Process in Evolution. Princeton University Press, Princeton.
Pedersen, E.A. and Widgren, M. (2011). Agriculture in Sweden, 800 BC – AD 1000. In The
Agrarian History of Sweden: 4000 BC to AD 2000, Myrdal, J. and Morell, M. (eds).
Nordic Academic Press, Lund.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Claudius’ Coin in the Forest – Niche Construction 219

Slotte, H. (2001). Harvesting of leaf-hay shaped the Swedish landscape. Landsc. Ecol., 16,
691–702.
von Stedingk, H. and Baudou, E. (2006). Capitalism in central Norrland, Sweden, during the
Iron Age. Curr. Sw. Archaeol., 14, 177–198.
Svensson, E. (2018). The scandinavian shieling – Between innovation and tradition. In
Historical Archaeologies of Transhumance across Europe, Costello, E. and Svensson, E.
(eds). Routledge, London.
Zeder, M.A. (2016). Domestication as a model system for niche construction theory. Evol.
Ecol., 30, 325–348.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
17

Recent History of Vegetation


Changes in the Arctic
Antoine BECKER-SCARPITTA1,2, Bastien PARISY1 and Tomas ROSLIN1,3
1
University of Helsinki, Finland
2
Institute of Botany of the Czech Academy of Science, Průhonice, Czech Republic
3
Swedish University of Agricultural Sciences, Uppsala, Sweden

17.1. Introduction

While human-induced global change is affecting all regions on Earth, it is in the


Arctic that its effects on ecological communities and native populations are
currently best visible (Post et al. 2019). There, the climate is warming at a rate of
nearly twice the global average, with a direct effect on permafrost, both sea and land
surface ice, the snow dynamic, and the functioning of the ecosystem (Box et al.
2019). Understanding and describing what is happening in the Arctic can thus help
inform future projections for the rest of the Earth. The Arctic region is also home to
several indigenous human populations, who – being, for example reindeer herders or
hunter-gatherers – depend on the local vegetation. A sudden change in vegetation
has direct consequences for their lifestyles, migration patterns and survival.

To reveal the effects of climate on ecosystems, the archives of historical ecology


– i.e. observations made over the past decades – provide an unparalleled resource. In
this chapter, we start with a brief description of the Arctic tundra biome, and then
summarize how historical data have been used to describe and understand recent
trends in Arctic tundra vegetation. In doing so, we focus on changes observed during
the last five decades, while refraining from digging into the paleoecological records.

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
222 Historical Ecology

17.2. The Arctic tundra biome

Stretching north from the latitudinal tree line, the Arctic region is characterized
by tundra vegetation – i.e. low-productivity rate vegetation mainly composed of
herbaceous plants, shrubs, bryophytes and lichens. In total, the Arctic tundra biome
spans nearly 7 million km2, and is characterized by harsh environmental conditions.
Winters are long and cold, and the growing season short, relatively cold, and
typically dry. Direct human activity in the area is generally low, with the exception
of herbivory pressure from reindeer herding.

Climate warming in the Arctic is proceeding at a rapid pace. While global


surface temperatures have increased by 2°C on average in recent decades, annual
mean temperatures in the Arctic region have risen by an average of 4°C – whereas
wintertime temperatures have increased by up to 7°C (Box et al. 2019). The faster
and greater change in winter, as compared to other seasons, has resulted in earlier
snowmelt, lower levels of ice and snow cover and major biological changes (Arctic
Report Card 2020 – https://www.arctic.noaa.gov/). At the same time, the Arctic
region has experienced a general increase in air humidity and precipitation –
although with high local variations. Moreover, under all climatic scenarios, the
Arctic will continue warming more rapidly than any other region on Earth (IPCC
2014).

The current pattern of warming is being imposed on a flora adapted to harsh


climatic conditions, with typical plant traits including short stature, waxy leaves and
dense leaf hairs. Tundra plants typically form a dense mat of vegetation, or compact,
cushion-like clusters, and their roots tend to grow along the soil surface, thereby
taking advantage of the warmest and best-drained layers of the soil. Species
diversity is low, and much of the biome is dominated by a relatively few species,
each of which is widely distributed across the biome (Walker et al. 2018). We can
distinguish five major growth forms among tundra plants, based on their physical,
ecological and biogeochemical characteristics. Described progressing from warmer
to colder conditions, they are: (i) low- and high-shrub, (ii) erect dwarf-shrubs,
(iii) prostrate dwarf-shrubs, (iv) forb-lichen-moss cushions and (v) graminoids and
forbs, which grow under the driest soil conditions (Raynolds et al. 2019).

17.3. The Arctic historical ecological archive

The historical ecological records amassed since the 20th century provide
valuable data for distinguishing the drivers of population and community responses
to global change (Vellend et al. 2013). This historical data is derived from old
satellite images or photographs, the long-term monitoring of permanent plots and
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Recent History of Vegetation Changes in the Arctic 223

semi-permanent vegetation plots, and repeated visits to previously surveyed sites.


Given the relatively short history of arctic ecology, we will mainly focus on records
from the 1960s and later.

17.3.1. Remote sensing over time

The remoteness of the Arctic biome imposes both logistical and financial
challenges for anyone wanting to conduct field research, especially across multiple
sites. Thus, remote sensing methods are of paramount importance for studying arctic
change at the landscape or regional scale (Beamish et al. 2020). Of these methods,
satellite-based imagery provides some of the most powerful testimonies of
ecological changes (Myers-Smith et al. 2020). Another important remote sensing
method is the use of historical aerial photographs. Such imagery provides
information that can complement satellite data, and allows us to see back in time to
the pre-satellite era. Satellite imagery can be used to quantify different types of
vegetation through their spectral reflectance signatures. The dominant metric used
over the last four decades is the Normalized Difference Vegetation Index (NDVI),
which is based on the reflectance signature of the vegetation. Since being developed
in the late 1980s, NDVI has been widely used as a proxy for measurement of
photosynthetic activity and vegetation productivity in the Arctic.

The use of NDVI is not without complications, since the spectral signature of
plants may be comingled with other abiotic factors at the landscape scale, and since
the interpretation of images is subject to influence from the methodological choices
made. Satellite images are sensitive to clouds and snow cover, as well as to water
retention and topography. From a methodological perspective, the resolution and
accuracy of the images may vary with the sensor type and satellite technology used,
with the algorithms applied to the images, and with the specific timing of the
observations. Overall, the very nature of the arctic biome offers several challenges to
studying it, since radiation enters at a low angle in the region, the growing season is
short, and landscapes are typically heterogeneous (Beamish et al. 2020).

17.3.2. Field-based records

17.3.2.1. Long-term monitoring of permanent plots


To work across multiple remote sites, arctic ecologists early on joined
forces in an effort called the International Tundra Experiment (ITEX;
https://www.gvsu.edu/itex/). Established in the 1990s, this initiative is based on a
standardized vegetation monitoring protocol that is designed to be implemented by
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
224 Historical Ecology

different researchers at different sites. A primary objective of ITEX has been to


manipulate temperature with open-top chambers (OTCs), to thereby test the effect of
warming on tundra plant phenology, growth, community composition and diversity.
To distinguish the specific effect of the warming treatments, the protocol also
includes the monitoring of control plots.

Spanning over 30 years of standardized plot-level studies, the ITEX data now
provide some of the largest, most comprehensive and longest time-series of
vegetation change in the arctic biome. Other local long-term monitoring initiatives
span different areas and time periods, such as the Greenland Ecological Monitoring
(GEM – https://data.g-e-m.dk/) program. In the nature reserves of the former USSR,
communities and seasonal events have been systematically recorded for decades,
and even sometimes for over a century (Ovaskainen et al. 2020).

17.3.2.2. Resurveys of semi-permanent vegetation plots


In the 1960s, the main objective of plant ecology was typically to classify
vegetation in relation to local environmental descriptors. Such sampling campaigns
were mainly aimed at a type of vegetation inventory called phytosociology (or
geobotany). A valuable legacy of these campaigns is the standardized protocol that
they developed. As their plots can oftentimes be relocated and resampled decades
later, ecologists can then infer what changes have occurred in the intervening period.
Although legacy studies in the Arctic have been limited to a relatively small number
of sites, they are of great value in testing hypotheses about the environmental drivers
of vegetation-level change (Callaghan et al. 2011).

17.3.2.3. Historic field observations and herbaria


The ecological records held in museums, in herbaria, and at research stations can
offer valuable sources of information on historical community composition,
phenology and population dynamics. While observations and experiments from the
Arctic tend to be clumped together in space and to be brief in duration, herbarium
specimens collected over centuries allow us to explore another dimension of
long-term ecological change (Lang et al. 2019). Over the last two decades, museums
across the world have initiated the digitization of their specimens. Now the resulting
open access data allows ecologists to resolve ecological questions. In the Arctic, the
combination of different datasets from herbarium specimens, photographs and field
observations have allowed novel descriptions of the changes taking place in tundra
plant phenology in response to warming (Panchen and Gorelick 2017). However,
unlike semi-permanent plots, which can be relocated with some confidence,
botanical descriptions of an area are rarely associated with exact locations. This
weakens the ecological conclusions that can be derived from this type of data.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Recent History of Vegetation Changes in the Arctic 225

17.4. Changes over time in tundra vegetation

In this section, we provide an overview of the recent changes in arctic tundra


vegetation revealed by the historical archives detailed above. During the last few
decades, we have seen three major dimensions of ecological change, namely,
changes in (1) vegetation productivity, (2) vegetation phenology and (3) plant
community structure, composition and diversity.

17.4.1. Changes in vegetation productivity

Large-scale analyses of satellite images indicate a general, long-term increase in


NDVI in the Arctic tundra region. This trend has been termed “Arctic greening”,
referring to an increase in plant productivity. Conversely, a decrease in NDVI is
referred to as “browning”, suggesting a decrease in plant productivity or vegetation
cover. The pattern of greening was first described over the period 1981–1991, and is
now considered one of the clearest continental-scale responses to recent
human-induced global warming. The underlying mechanisms of greening appear to
be: (i) an increase in existing shrub patches, commonly refer to as expansion, (ii) an
increase in plant productivity, including an increase in plant height and leaf area and
(iii) a shift in species distribution, mainly due to colonization (Myers-Smith et al.
2011).

Although a growing body of evidence has linked human-induced warming over


the past 40 years to the decadal change in vegetation productivity, the direction and
intensity of this change have been far from uniform across the Arctic (Huang et al.
2017; Berner et al. 2020). Over the past two decades, several sources indicate a
reversal in the direction of NDVI change in the Arctic biome, from greening to
browning. Indeed, since the 2000s the global increase in NDVI has slowed. In
north-eastern Eurasia, and northern and northwestern America, it appears to have
even decreased. Several active research programs are currently exploring regional
differences in patterns of productivity change, attempting to explain the causes. In
particular, increased snow cover in the winter and a late spring snowmelt have been
suggested as potential explanations (Bhatt et al. 2013). On top of these directional
trends in NDVI, topographic position, the composition of the vegetation (e.g. shrub
cover), water balance and snowmelt dynamics appear to generate additional
geographic variation in plant productivity (Reichle et al. 2018).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
226 Historical Ecology

17.4.2. Changes in vegetation phenology

Phenology refers to the timing of recurrent life cycle events in living organisms.
Plant phenology includes, for example, the timing of flowering, seed maturation,
bud break or senescence. For most plant species, phenological events are often
closely linked to temperature or photoperiod. Thus, in the context of global climate
change, changes in phenology are a key indicator of a species response to warming
(Thackeray et al. 2016). In recent decades, remarkable changes have been observed
in the phenology of plants in the Arctic tundra biome. These observations derive
from long-term plot-level monitoring and experiments (Prevéy et al. 2019), from
remote sensing (Beamish et al. 2020), and from herbarium specimens, historical
photographs and field observations (Panchen and Gorelick 2017).

17.4.2.1. Changes in timing


Beyond providing a measure of primary productivity, data on NDVI provide a
useful measure of the time of the season when this productivity occurs, and thus of
phenology. Time series of satellite-derived NDVI have been widely used to detect
an earlier start and a later end of the growing season, and thus of a lengthening
growing season (Zeng et al 2013). Warming experiments conducted at the ITEX
sites suggest an advancement in the timing of both leaf emergence and flowering
time, whereas they show no change in or a slight delay of leaf senescence
(Bjorkman et al. 2020). However, the direction and magnitude of phenological
changes vary considerably between biomes, and within a biome between different
species (Roslin et al. 2021).

17.4.2.2. Shortening flowering seasons


In an Arctic-wide synthesis, Prevéy et al. (2019) used records from long-term
monitoring and warming experiments to test for a relationship between the
temperature sensitivity and phenological niches of 253 tundra species. What these
authors detected was a greater change in late-flowering species than in
early-flowering species. Thus, they concluded that the community-wide flowering
season is contracting in response to climate warming, and that this contraction is
due to an advance in the flowering date of late-flowering species – without
corresponding changes in early-flowering species. Such a shortening of the
flowering season is consistent with other local studies in the Arctic region (Høye
et al. 2013). Although increasing air temperature has a direct effect on local
phenology, the snow cover modulates the effect of air temperature on conditions in
the soil. Thus, snow dynamics modify the relationship between temperatures and
realized phenology (Kelsey et al. 2021).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Recent History of Vegetation Changes in the Arctic 227

17.4.2.3. Cascading effects on trophic interactions


A shorter flowering season will directly limit resource availability for higher
trophic levels (e.g. pollinators or herbivores), thereby affecting food web structure
(Hegland et al. 2009). Since the growing season in the Arctic region is short to begin
with, different trophic levels are generally well synchronized in time. Contrasting
shifts in phenology between plants and higher trophic levels might then lead to a
phenological mismatch (Høye et al. 2013). In a resource-limited environment such
as the Arctic tundra, we may also see competition for pollinators among plants.
Climate change can thereby result in a negative feedback loop from pollinators back
to the plant community (Schmidt et al. 2017).

17.4.3. Changes in plant community structure, composition and


diversity

Environmental changes in the Arctic have resulted in shifts in community


composition, in species’ abundance and distribution, and in plant functional traits
(Nabe-Nielsen et al. 2017; Bjorkman et al. 2018a). As remote sensing methods have
yet to reach the accuracy needed to reliably detect such small-scale changes,
relevant historical data are therefore mainly limited to field observations. In this
context, the ITEX site network provides the most extensive and longest-term
monitoring of community composition in the tundra biome. A synthesis using data
from 46 individual ITEX control plots (1980–2010) revealed a relationship between
temperature warming and changes in community composition, despite high
variability between sites (Elmendorf et al. 2012). What was specifically found was
an increase in the size of herbaceous plants, in the organic layer of the soil, and in
the abundance of shrubs and herbaceous plants. At the same time, the area of bare
soil was reduced.

Complementing the findings from the ITEX sites, the “Back to the Future”
research program has generated repeated observations of selected local
communities. Similar patterns emerged across the individual sites that were
revisited: a general sensitivity of tundra vegetation to warming and to changing
precipitation regimes, an increase in shrub abundance and a shift in shrub
distribution. Equally important is the finding that patterns differ considerably
between sites, highlighting a strong context-dependence in the responses of tundra
plant communities to warming (Callaghan et al. 2011).

17.4.3.1. Change in functional traits


Until recently, the exploration of functional traits among Arctic plants was
mostly restricted to coarse functional grouping based on growth forms such as
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
228 Historical Ecology

shrubs, forbs, graminoids, mosses and lichens. Such crude classification of plants by
their appearances is hardly linkable to key ecosystem functions. Yet, in 2018, the
publication of the database generated by the Tundra Trait Team (TTT) marked a
milestone in the history of Arctic trait-based ecology. This database builds on field
measurement of 18 key functional traits for 978 tundra species, which traits can be
directly linked to key ecosystem functions and assembly rules (Bjorkman et al.
2018b). This conquering of the functional dimension allows ecologists to test
hypotheses related to how current change is affecting community and ecosystem
functioning at a biogeographical scale.

By using the plant Tundra Trait database, Bjorkman et al. (2018a) quantified
changes in intraspecific variability and community-level variability over time.
Consistent with several studies showing increases in plant productivity and growth
rate (see above), the trait-based approach reveals an increase in weighted community
height across the biome over the past three decades, while other traits (e.g. leaf
traits) appear not to have changed over time. The observed change in plant
community height seems partly related to changes in abiotic conditions and to
changes in species composition, rather than to changes in species abundance. In
other words, the observed change in vegetation height reflects the range expansion
of taller species towards higher latitudes and higher elevations in response to
warming.

17.4.3.2. Shrubification
A major change in the functional composition or growth form of tundra plants is
the expansion of shrubs. Using historical oblique aerial photographic tracking from
1945, Tape et al. (2006) demonstrated that shrubs have expanded over a large area in
Alaska. Another example of the use of historical aerial photography is the study by
Frost and Epstein (2014), who took advantage of the declassification of
high-resolution photographs of the former USSR, taken in the 1960s by US
intelligence, to demonstrate that shrubs and trees have expanded in the tundra
ecotone since that time.

As mentioned above, a general greening of the tundra (see section 17.4.1) has
been attributed to shrub expansion. In terms of community structure, shrubification
is mainly attributed to a positive response to warming and to changing precipitation
regimes by genera such as Betula, Salix and Alnus, and also by Empetrum, Cassiope
and others. The trends observed by remote sensing are supported by in situ
observations of changes in shrub abundance, of shifts in community composition
and in species distribution, and of increases in annual growth rings and in plant size
(Myers-Smith et al. 2019). The expansion of shrubs across the tundra has direct
consequences for ecosystem functioning, as it may inter alia decrease albedo, thus
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Recent History of Vegetation Changes in the Arctic 229

resulting in altered snow accumulation and melting dynamics, changes in carbon


cycling by increasing photosynthetic activity, and changes in the thickness of the
organic layer. However, shrubification may also have direct consequences for biotic
communities, causing changes in understory species composition through changes in
canopy conditions (i.e. in light, temperature and humidity), and changes in the
herbivory regime and/or in microbial and fungal activity (Myers-Smith et al. 2011).

17.5. Synthesis and perspectives

A growing body of evidence links human-induced warming over the past four
decades to changes in the species composition, productivity and structure of arctic
vegetation. However, the direction and trends of this change vary across space, time
and taxa (Huang et al. 2017). Using three major ecological patterns of change –
productivity, phenology and community composition – we have demonstrated the
value of different historical ecological archives for identifying and understanding
these changes. Our synthesis highlights that imprints are most clearly detected when
complementary data of multiple types are considered. Yet, importantly, such
comparisons also expose conflicts in the evidence provided by different types of
historical data. As an example of consistent trends exposed by multiple types of
evidence, the greening trend observed over the tundra biome – and more recently,
the browning observed within given regions – was suggested by remote sensing but
later verified through in situ observations of shrubification in the Arctic landscape.
A large body of evidence now supports an increase in plant height and shrub
abundance, in the face of limited temporal change in vegetation composition. As an
example of conflicting evidence, remote sensing methods have pointed to an earlier
start and later end of the growing season, resulting in a longer growing season. By
comparison, long-term in situ monitoring of phenology has revealed a general
shortening of the flowering season.

Importantly, the major trends of vegetation change detected across the Arctic are
almost invariably associated with a high variability in the direction and magnitude of
change across space, over time and among species. This points to a strong context-
dependence in how tundra vegetation responds to climate change. The spatial
heterogeneity of ecological changes in the tundra vegetation underlines the need for
widespread observations conducted under variable conditions. One major challenge
then is the highly uneven spatial coverage of the current ecological archive. There is
thus an urgent need to extend current research to new sites, thereby improving our
understanding of arctic complexity.

As discussed elsewhere in this book, and specifically for the arctic region in this
chapter, historical records provide a crucial avenue towards detecting long-term
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
230 Historical Ecology

ecological change. Every arctic study, and in fact every observation, has the
potential to contribute to the historical archives of the decades to come. It is
therefore the responsibility of every contemporary scientist to structure their data in
standard formats and to store them in openly accessible databases – thereby
benefitting their future peers and colleagues.

17.6. References

Arctic Report Card (2020). Arctic Program [Online]. Available at: https://www.arctic.noaa.gov/
[Accessed March 2021].
Beamish, A. et al. (2020). Recent trends and remaining challenges for optical remote sensing
of Arctic tundra vegetation: A review and outlook. Remote Sensing of Environment, 246,
111872 [Online]. Available at: https://doi.org/10.1016/j.rse.2020.111872.
Berner, L.T. et al. (2020). Summer warming explains widespread but not uniform greening in
the Arctic tundra biome. Nature Communications, 11(1), 4621. doi: 10.1038/s41467-020-
18479-5.
Bhatt, U. et al. (2013). Recent declines in warming and vegetation greening trends over
pan-arctic tundra. Remote Sensing, 5(9), 4229–4254. doi: 10.3390/rs5094229.
Bjorkman, A.D. et al. (2018a). Plant functional trait change across a warming tundra biome.
Nature, 562(7725), 57–62. doi: 10.1038/s41586-018-0563-7.
Bjorkman, A.D. et al. (2018b). Tundra trait team: A database of plant traits spanning the
tundra biome. Global Ecology and Biogeography, 27(12), 1402–1411. doi:
10.1111/geb.12821.
Bjorkman, A.D. et al. (2020). Status and trends in arctic vegetation: Evidence from
experimental warming and long-term monitoring. Ambio, 49(3), 678–692. doi: 10.1007/
s13280-019-01161-6.
Box, J.E. et al. (2019). Key indicators of arctic climate change: 1971–2017. Environmental
Research Letters, 14(4), 045010. doi: 10.1088/1748-9326/aafc1b.
Callaghan, T.V. et al. (2011). Multi-decadal changes in tundra environments and ecosystems:
Synthesis of the international polar year-back to the future project (IPY-BTF). Ambio,
40(6), 705–716. doi: 10.1007/s13280-011-0179-8.
Elmendorf, S.C. et al. (2012). Plot-scale evidence of tundra vegetation change and links to
recent summer warming. Nature Climate Change, 2(6), 453–457. doi: 10.1038/
nclimate1465.
Frost, G.V. and Epstein, H.E. (2014). Tall shrub and tree expansion in Siberian tundra
ecotones since the 1960s. Global Change Biology, 20(4), 1264–1277. doi: 10.1111/gcb.
12406.
GEM (2021). The Greenland Ecosystem Monitoring (GEM) Database [Online]. Available at:
https://data.g-e-m.dk/ [Accessed March 2021].
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Recent History of Vegetation Changes in the Arctic 231

Hegland, S.J. et al. (2009). How does climate warming affect plant-pollinator interactions?
Ecology Letters, 12(2), 184–195. doi: 10.1111/j.1461-0248.2008.01269.x.
Høye, T.T. et al. (2013). Shorter flowering seasons and declining abundance of flower visitors
in a warmer Arctic. Nature Climate Change, 3(8), 759–763. doi: 10.1038/nclimate1909.
Huang, M. et al. (2017). Velocity of change in vegetation productivity over northern high
latitudes. Nature Ecology & Evolution, 1(11), 1649–1654. doi: 10.1038/s41559-017-
0328-y.
IPCC (2014). Climate Change 2014: Synthesis Report. Contribution of Working Groups I, II
and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate
Change. Geneva [Online]. Available at: https://www.ipcc.ch/report/ar5/syr/.
ITEX (2021). International Tundra Experiment (ITEX) [Online]. Available at: https://www.
gvsu.edu/itex/ [Accessed March 2021].
Kelsey, K.C. et al. (2021). Winter snow and spring temperature have differential effects on
vegetation phenology and productivity across Arctic plant communities. Global Change
Biology, 27(8), 1572–1586. doi: 10.1111/gcb.15505.
Lang, P.L.M. et al. (2019). Using herbaria to study global environmental change. New
Phytologist, 221(1), 110–122. doi: 10.1111/nph.15401.
Myers-Smith, I.H. et al. (2011). Shrub expansion in tundra ecosystems: Dynamics, impacts
and research priorities. Environmental Research Letters, 6(4), 045509. doi: 10.1088/1748-
9326/6/4/045509.
Myers-Smith, I.H. et al. (2019). Eighteen years of ecological monitoring reveals multiple
lines of evidence for tundra vegetation change. Ecological Monographs, 89(2), e01351.
doi: 10.1002/ecm.1351.
Myers-Smith, I.H. et al. (2020). Complexity revealed in the greening of the Arctic. Nature
Climate Change, 10(2), 106–117. doi: 10.1038/s41558-019-0688-1.
Nabe-Nielsen, J. et al. (2017). Plant community composition and species richness in the High
Arctic tundra: From the present to the future. Ecology and Evolution, 7(23), 10233–10242.
doi: 10.1002/ece3.3496.
Ovaskainen, O. et al. (2020). Chronicles of nature calendar, a long-term and large-scale
multitaxon database on phenology. Scientific Data, 7(1), 47. doi: 10.1038/s41597-020-
0376-z.
Panchen, Z.A. and Gorelick, R. (2017). Prediction of Arctic plant phenological sensitivity to
climate change from historical records. Ecology and Evolution, 7(5), 1325–1338. doi:
10.1002/ece3.2702.
Post, E. et al. (2019). The polar regions in a 2°C warmer world. Science Advances, 5(12),
1–12. doi: 10.1126/sciadv.aaw9883.
Prevéy, J.S. et al. (2019). Warming shortens flowering seasons of tundra plant communities.
Nature Ecology & Evolution, 3(1), 45–52. doi: 10.1038/s41559-018-0745-6.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
232 Historical Ecology

Raynolds, M.K. et al. (2019). A raster version of the circumpolar arctic vegetation map
(CAVM). Remote Sensing of Environment, 232(Oct), 111297. doi: 10.1016/j.rse.2019.
111297.
Reichle, L.M. et al. (2018). Spatial heterogeneity of the temporal dynamics of arctic tundra
vegetation. Geophysical Research Letters, 45(17), 9206–9215. doi: 10.1029/2018GL0
78820.
Roslin, T. et al. (2021). Phenological shifts of abiotic events, producers and consumers across
a continent. Nature Climate Change, 11(3), 241–248. doi: 10.1038/s41558-020-00967-7.
Schmidt, N.M. et al. (2017). Interaction webs in arctic ecosystems: Determinants of arctic
change? Ambio, 46(S1), 12–25. doi: 10.1007/s13280-016-0862-x.
Tape, K. et al. (2006). The evidence for shrub expansion in Northern Alaska and the Pan-
Arctic. Global Change Biology, 12(4), 686–702. doi: 10.1111/j.1365-2486.2006.01128.x.
Thackeray, S.J. et al. (2016). Phenological sensitivity to climate across taxa and trophic
levels. Nature, 535(7611), 241–245. doi: 10.1038/nature18608.
Vellend, M. et al. (2013). Historical ecology: Using unconventional data sources to test for
effects of global environmental change. American Journal of Botany, 100(7), 1294–1305.
doi: 10.3732/ajb.1200503.
Walker, D.A. et al. (2018). Circumpolar arctic vegetation classification. Phytocoenologia,
48(2), 181–201. doi: 10.1127/phyto/2017/0192.
Zeng, H. et al. (2013). Shifts in arctic phenology in response to climate and anthropogenic
factors as detected from multiple satellite time series. Environmental Research Letters,
8(3), 035036. doi: 10.1088/1748-9326/8/3/035036.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
18

Reconstructing the Impact of


Humans on Aotearoa New
Zealand’s Biodiversity
Nicolas J. RAWLENCE1, Alexander J.F. VERRY1,2, Karen GREIG1,
Justin J. MAXWELL3, Lara D. SHEPHERD4 and Richard WALTER1
1
University of Otago, Dunedin, New Zealand
2
University of Toulouse, France
3
Sunrise Archaeology, Mangonui, New Zealand
4
Museum of New Zealand Te Papa Tongarewa, Wellington, New Zealand

18.1. Introduction

Aotearoa New Zealand’s (NZ) biodiversity has been described as “… close as


we will get to the opportunity to study life on another planet” (Diamond 1990). As
an isolated archipelago, its biodiversity has been shaped by both tens of millions of
years of isolation and frequent long-distance dispersal. A dynamic geological history –
including partial Oligocene inundation, tectonic uplift, Pleistocene glacial-interglacial
cycles and volcanism – has also played a large role in the evolution of NZ’s biota.
At the time of Polynesian arrival, NZ’s vertebrate fauna was dominated by birds and
reptiles, which exhibited a range of island adaptations including gigantism,
flightlessness and K-selected breeding strategies (Worthy and Holdaway 2002).
There were at least 217 avian species breeding in the NZ ecological region at the
time of human arrival, ranging in size from giant moa (Dinornithiformes) and the
world’s largest eagle (Haast’s eagle, Aquila moorei) to among the smallest of birds,
the Acanthissittid wrens (Worthy et al. 2017). The herpetofauna was dominated by
Eugongylinae skinks, Diplodactylid geckos, tuatara (the last surviving member of

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
234 Historical Ecology

the Rhynchocephalia family) and Leiopelmatid frogs (Worthy and Holdaway 2002).
NZ also had a diverse marine mammal assemblage which included multiple species
of cetaceans and pinnipeds (Seersholm et al. 2018).

Paleoecological and archaeological evidence suggests Polynesians colonized NZ


in the late 13th century CE (Wilmshurst et al. 2008) at a time of relative climatic
stability (Rawlence et al. 2016a), followed by European settlement from 1796 CE.
In addition to hunting and environmental modification (Perry et al. 2014a, 2014b),
humans also introduced exotic predators (e.g. rats, dogs, mustelids), plants and
changed the genetic landscape through translocations and plant cultivation
(Shepherd et al. 2018). By the early 20th century, the NZ ecosystem had
fundamentally changed with the extinction of ~50% of its biodiversity, biological
turnover events, range contractions, population bottlenecks and significant changes
in indigenous forest cover (Waters et al. 2017; Rawlence et al. 2019).

Our understanding of these dramatic changes has taken a step-change with


multidisciplinary research incorporating morphological analysis of the
paleontological and archaeological record, in combination with radiocarbon-dating,
stable dietary isotopes and ancient DNA. This chapter focuses on the anthropogenic
impacts on NZ biodiversity up until the early 20th century, from an archaeological,
paleoecological and evolutionary perspective.

18.2. Archaeological evidence for anthropogenic impact in New


Zealand

NZ was settled at the last phase of a maritime expansion across the Pacific. As
settlers moved east across the Pacific Ocean, islands became increasingly small and
isolated. There was a natural biodiversity decline to the east, although most islands
supported abundant bird populations (Steadman 2006). These species evolved with
few terrestrial predators and the finely balanced ecosystems were disrupted when
humans arrived with domestic pigs, fowl, dogs, rats and plants, and cleared land for
agriculture. Naïve avifauna, particularly flightless and ground nesting birds, were
susceptible to hunting. What followed was a wave of extirpation and extinction
recorded in the archaeological records of many islands (Steadman 2006).

This pattern of human impact following the establishment of horticulture and


hunting, which led to rapid faunal extinctions, was repeated across Polynesia and
eventually NZ. The first settlers came to NZ from East Polynesia ~1300 CE
(Wilmshurst et al. 2008) from “Hawaiki”, which was probably located in a zone
encompassing the Cook, Austral and Society Islands. Archaeological evidence
throughout the “Hawaiki Zone” shows that islands were linked into an interaction
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reconstructing the Impact of Humans on Aotearoa New Zealand’s Biodiversity 235

sphere; a zone of regular inter-island voyaging and exchange (Walter et al. 2017).
Colonization and the subsequent pattern of networked mobility were enabled by
long-distance voyaging craft, navigation skills and a portable economy – the ability
to transform natural landscapes into cultural ones. NZ was settled out of this
dynamic zone of interaction.

Arriving in NZ, the first settlers adopted an effective strategy for systematic
exploration of coastlines, river systems and resource bases. Genetic evidence
suggests a large founder population was required to account for current levels of
diversity, implying multiple settlement voyages over a short period (Walter et al.
2017). By the mid-1300s CE, NZ had been explored; all major resources had been
discovered and were being widely traded in coastal exchange networks (Walter et al.
2017). Settlement spread rapidly and a pattern emerged involving dispersed coastal
occupation, with clusters of sites around harbors, and larger villages at river mouths.

The first settlers’ strategy was to reproduce settlement patterns, economics and
lifestyles of their tropical homeland; characterized by high levels of mobility
facilitated by ocean-going canoes, long-distance exchange of high-quality resources,
the maintenance of social relationships, and the significance of harbors, river mouths
and estuaries for settlement. As in “Hawaiki”, populations were dispersed (but
connected) and like their Pacific ancestors, settlers took up hunting to supplement
their horticultural and fishing economy for a generation or two.

Moa were the most important prey species and their significance in the Early
Māori (pre-1500 CE) economy is well reflected in the archaeological record.
Numerous large “moa-hunter” village sites have been excavated around river
mouths in NZ, containing the remains of hundreds of moa that were hunted in the
hinterlands and brought to the villages for butchering and consumption. The rich
material culture and ornate burial practices of these early communities reflect a high
level of wealth and security. In addition to large villages, small seasonal hunting
sites and specialized processing sites have been identified, containing the butchered
and cooked remains of one to two moa, plus associated stone tools (Anderson 1989).
Discussion of moa exploitation revolves around the interpretation of bone
assemblages from early sites. There is abundant evidence, however, that hunters
were also engaged in high levels of egg predation. Most sites with moa bone also
contain eggshell, often in large quantities, and a few early sites have produced whole
moa eggs (Anderson 1989). As in Polynesia, NZ avifauna evolved in the absence of
mammalian predators including humans. Of the 131 known endemic species present
in NZ at the time of Polynesian settlement, at least 40 became extinct prior to
European arrival at a time of relative climatic stability (Duncan and Blackburn
2004) as a result of different birding strategies adopted by early Māori.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
236 Historical Ecology

There is little compelling evidence that non-moa bird hunting in NZ was ever
intensive, or involved specialized, taxon-specific hunting strategies. It appears to
have been largely opportunistic (Anderson 1989) and taxon-specific impacts as a
result of hunting can be reasonably predicted on the basis of habitat, flight ability,
body size and nesting behavior (Duncan and Blackburn 2004). The archaeological
record documents the scale, range and impact of hunting, but little is understood
about hunting practices, strategies and the actual process by which targeted faunas
were affected. Some models of hunting behaviors leading to moa extinction have
been proposed, including Anderson’s (1989) discussion of rolling wave, blitzkrieg
or serial overkill scenarios. But these, and other models of overkill, assume a
lengthy period of hunting, with colonization up to 1,000 years ago. With refinements
in radiocarbon-dating it is now apparent that settlement of NZ occurred ~1300 CE
(Wilmshurst et al. 2008). Furthermore, Perry et al. (2014a) demonstrated that moa
were extinct no more than 150 years later, and perhaps less. The problem of moa
extinction rates has been addressed in genetic models (e.g. Allentoft et al. 2014), but
there is almost nothing concrete known about moa hunting practices from the
archaeological record. So far, no hunting tools have been identified and there is no
evidence on bones (e.g. impact trauma or dog tooth marks) that might indicate a
killing method. Whatever answer, in terms of per-annum kill rate, that ecological
modeling might predict for these K-selected birds, this is only a small part of the
puzzle. The bigger question revolves around behavioral aspects of predation – what
were Māori actually doing to have such a rapid impact on moa. The situation is no
less enigmatic for the problem of pinniped hunting.

NZ’s coastline had abundant fur seal (Arctocephalus forsteri) and sea lion
(Phocarctos hookeri) colonies on both main islands (Smith 1989). By European
contact, fur seals were restricted to the southwestern South Island, and sea lions and
southern elephant seals (Mirounga leonina) were extirpated from mainland NZ
(Rawlence et al. 2016a). Pinnipeds have been reported from early Māori sites around
NZ but are absent from North Island sites by 1500 CE and decline sharply after that
in the south with cryptic biological turnover events (Waters et al. 2017). Smith
(1989) reconstructed two strategies adopted by Māori hunters: regular seasonal
cropping of breeding colonies and opportunistic killing of beached individuals. The
former was the most economically important and had the highest impact, the latter
was less frequent but occurred more widely throughout NZ.

Unlike moa hunting, there are very few sites with abundant pinniped bone and
thus little direct evidence for large-scale hunting. Instead, the evidence most
frequently cited that Māori heavily impacted pinnipeds is the reduction in their
range. The presence of breeding colonies in northern NZ at the time of human
arrival (Smith 1989) implies a significant range contraction, as a result of early
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reconstructing the Impact of Humans on Aotearoa New Zealand’s Biodiversity 237

Māori predation. In fact, it implies a greater range contraction than that caused by
18th century CE industrial sealing. So far, predation on that scale is not directly
supported by archaeological evidence. However, genetic modeling indicates sea lion
extirpation is possible even with low but sustained levels of hunting (e.g. < 0.5
individuals/person/year; Rawlence et al. 2016a).

18.3. Paleovegetation change in pre- and post-European contact New


Zealand

Environmental modification was an important part of the successful colonization


of Polynesia, and fire was the principal tool to modify vegetation communities. In
NZ, it has been argued that large-scale vegetation change occurred rapidly. The
range of hypothesized drivers for change include improving the landscape for
cultivation, collection or encouragement of wild crops, or to remove dense forest
and improve travel or resource access (McGlone et al. 2005). Vegetation
modifications may have been unintentional, resulting from other activities.

Palynology and charcoal analysis have been the main methods used to study
vegetation change through time. Most pollen records are used to investigate natural
processes of long-term, multi-millennial paleoenvironmental change, addressing
various themes especially climate change. These studies can also inform on fire
regimes, both locally and nationally; changes in sedimentation regimes; and the
timing and impact of human arrival, although anthropogenic changes are often only
a small component of this type of research.

It has been estimated that 85% of NZ’s landmass was covered in forest before
humans arrived (Perry et al. 2014b). Native forests varied greatly due to differences
in climate and topography across NZ. While forests were diverse, they generally
shared the following characteristics during the Holocene: low fire frequency, a
majority of endemic species not adapted to fire and, as a result, a majority of
ecosystems highly vulnerable to fire (Perry et al. 2014b).

Polynesian arrival coincided with a major change in fire frequency and intensity.
The initial burning period (IBP) was widespread and occurred within a few centuries
of human arrival (Perry et al. 2014b). The combination of forest species not adapted
to fire, human-lit fires and vegetation changes which occurred after firing all
increased flammability of these environments. This led to long-term environmental
modifications of vast areas, with loss during the IBP of 40% of the existing forests
and most of the lowland and montane dry forests which were the most fire
susceptible (Perry et al. 2014b).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
238 Historical Ecology

A plant that is admirably adapted to fire is the bracken fern (Pteridium


esculentum). Bracken recolonizes fire-affected areas, has a high reproductive rate
and survives repeated firing. It is not surprising then that after human arrival,
bracken fern became one of the most abundant species in pollen records and,
combined with charcoal, is often used as a proxy for pre-European anthropogenic
environmental modification. Bracken fern also produces large, edible rhizomes,
which were an important carbohydrate in the Māori diet. Edible bracken was mainly
recovered from deep, fertile soils and the presence of bracken in other locations has
been hypothesized to represent environmental management to improve traveling
conditions (McGlone et al. 2005).

Investigations of NZ paleovegetation, primarily designed to study pre-European


contact human activity are less common. A paleobotanical study of
human-generated vegetation change on Rēkohu (Chatham Island, one of NZ’s
offshore islands), which was settled by Moriori ~1450 CE, combined traditional
anthropological and anthracological (charcoal) datasets, enhancing our
understanding of human impact on small islands. Vegetation changes were rapid and
long-lasting. The Rēkohu pollen and charcoal record indicated that soon after human
arrival, fires resulted in long-term loss of much of the forest and a significant
increase in bracken fern. However, ethnographic, archaeological and anthracological
data, combined with palynology, indicated that coastal forests survived and were
heavily managed in a form of arboriculture (Maxwell et al. 2016). The primary
species in these managed forests was kōpi/karaka (Corynocarpus laevigatus), which
was translocated from the NZ mainland. Karaka pollen preserves poorly and is
rarely noted in pollen records, and is probably highly underrepresented (Maxwell
and Tromp 2016). Palynological records alone therefore could not have provided a
reliable representation of this species in the past, nor could they have contributed to
our understanding of its relative abundance or its importance to Moriori subsistence.

From 1769 CE, European arrival initiated what would be a second wave of
large-scale vegetation change in NZ. There was further forest loss and substantial
changes occurred when new food plants were introduced. The earliest industries in
post-European contact NZ centered on the harvest of natural resources from sea and
land. On land, forests were harvested from the 1820s CE to supply timber spars for the
British Navy, from the 1830s CE to supply cut timber to Australia and beyond and, as
the population of NZ increased, to supply building timber locally (King 2003).

From the 1840s CE onward, there was also widespread loss of forest for
agriculture. It is estimated that most of the forest loss (~90%) occurred early in the
European period, between 1830 and 1873 CE, including the widespread loss of
old-growth forests (Perry et al. 2014b). Fires that occurred at this time were large
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reconstructing the Impact of Humans on Aotearoa New Zealand’s Biodiversity 239

and often not intentionally set, as accounts often cite a combination of drought
conditions and specific ignition sources. Today, less than 25% of the land area of
NZ is forested.

18.4. Utilizing Aotearoa’s natural resources: Māori cultivation and


translocation of flora and fauna

As Polynesians colonized the Pacific, they took useful plants and animals with
them. At least some of these were introduced to NZ during settlement, for example,
kūmara (Ipomoea batatas) and kurī, the Polynesian dog (Canis familiaris).
However, many of the plants, which were important food staples in the tropical
Pacific, failed to establish or only flourished in warmer parts of NZ. Māori and
Moriori brought into cultivation a number of indigenous plant species, and it is also
suspected that they translocated several mollusks and fish far outside their natural
distribution to secure a more predictable resource supply (e.g. Maxwell and Tromp
2016; Ross et al. 2018). Given its late settlement, NZ offers a unique opportunity to
examine the early stages of cultivation and domestication (e.g. what stage in the
domestication process is genetic diversity lost), in conjunction with Māori oral
histories, which complement western scientific methods and provide a more
complete picture of past translocations.

There is evidence of selection for morphological characters in plants. Some


karaka trees growing near former Māori habitation sites have larger fruit compared
with wild trees (Maxwell and Tromp 2016). There was selection of a number of
cultivars of tī kōuka/cabbage tree (Cordyline australis), whose roots and stems were
used as a carbohydrate source. The main tī kōuka cultivar in the North Island was tī
para, a sterile form that could be easily propagated vegetatively. A number of plants
were grown for their fiber; the most prized of these was harakeke/NZ flax
(Phormium tenax). At least 60 cultivars of harakeke were developed for different
fiber characteristics (Wehi and Clarkson 2007).

Māori also greatly extended the distributions of some plants and animals through
translocations to new regions. Some movements were recorded in myth and oral
histories (see below), but others have been inferred based on unusual species
distributions including occurrence around sites of former Māori occupation. Powell
(1938) suggested that several flax snail (Placostylus spp.) populations around past
Māori occupation sites may have resulted from translocation. Karaka trees were
introduced outside their natural range in northern North Island into southern North
Island and northern South Island, as well as to the distant Kermadec and Chatham
Islands (800 km and 650 km from the mainland, respectively). This is supported by
oral histories about the origins of karaka groves (Atherton et al. 2015). There are
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
240 Historical Ecology

myths relating to some of the founding ancestors, that talk about the seeding of
rivers and lakes with indigenous species of fish and invertebrates, including kōaro
(Galaxias brevipinnis), kōura (Paranephrops planifrons) and kākahi (Echyridella
spp.; McDowall 2011).

Genetic data has been used to examine past Māori translocations. High levels of
genetic structuring within natural populations of rengarenga (Arthropodium
cirratum), an endemic plant with edible rhizomes, allowed inference that it had been
brought into cultivation at least three times from different source populations
(Shepherd et al. 2018). Similarly, genetic relationships between populations of
tītīrangi (Veronica speciosa) indicate that the sole extant South Island population
probably derived from human-mediated dispersal from one of the most distant
natural populations (Armstrong and de Lange 2005). Genetic data from karaka and
the surf clam toheroa (Paphies ventricosa) revealed low genetic diversity and a lack
of structuring in natural populations resulting in the source(s) of species
translocations being unclear (Atherton et al. 2015; Ross et al. 2018).
High-throughput sequencing methods may provide greater power for distinguishing
between such populations but have been little used for reconstructing translocation
histories in NZ to date. Other challenges for reconstructing Māori translocations
include distinguishing between natural and cultivated populations in species with
wide natural distributions and the impact of hybridization and European
translocations on genetic patterns. Techniques such as palynology, anthrocology,
radiocarbon-dating and paleoenvironmental DNA sampling of ancient sediments can
also inform about translocations, especially in determining the timing of
introductions (Maxwell et al. 2016).

Translocation of plants and animals by Māori contributed to the modification of


NZ. Such introductions led to changes in the structure of ecosystems, many of which
are not yet fully understood. There is the potential for translocated species to
become invasive in their new habitats (e.g. karaka). Dealing with such invasive
species can be challenging because the control of their invasiveness must be
balanced with preserving culturally significant plantings.

18.5. Evolutionary consequences of Polynesian and European arrival

Polynesian arrival in NZ resulted in the extinction of ~50% of animal species


present at the time, including at least 54 avian species, three frogs and one skink
(Worthy et al. 2017). Unique lineages of conspecific taxa also went extinct (e.g.
Rawlence et al. 2016a). This was achieved through hunting, environmental
modification and predation from kiore, the Pacific rat (Rattus exulans), and kurī
(Greig and Rawlence 2021). One of the most well-studied extinctions is that of the
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reconstructing the Impact of Humans on Aotearoa New Zealand’s Biodiversity 241

moa. With stable (or increasing) population sizes when humans arrived (Allentoft
et al. 2014), all nine species were extinct within 150 years (Perry et al. 2014a).
Extensive radiocarbon-dating of archaeological sites also shows that several other
taxa went extinct during this period. This includes Waitaha penguin (Megadyptes
waitaha) and the prehistoric NZ, and Chatham Islands lineages of sea lion, which
are only known from subfossil, and early archaeological sites (Rawlence et al.
2016a; Waters et al. 2017).

European arrival resulted in an additional wave of extinctions (~30 avian species


to date including huia Heteralocha acutirostris and laughing owl Ninox albifaces) as
new mammalian predators were introduced (e.g. mustelids, ship (R. rattus) and
Norway (R. norvegicus) rats), alongside further environmental modification. This
culminated in the arrival of ship rats on Big South Cape Island in the 1960s CE,
which resulted in the extinction of Stead’s bush wren (Xenicus longipes variabilis),
South Island snipe (Coenocorypha iredalei) and the greater short-tailed bat
(Mystacina robusta). One posterchild species of European impact was the relictual
Lyall’s wren (previously Stephen’s Island wren; Traversia lyalli), which was
extirpated from mainland NZ no doubt as a result of predation by kiore (Worthy and
Holdaway 2002). Subsequent predation by feral cats (Felis catus) rendered the bird
extinct by 1895 CE. Anthropogenic pressures facilitated by Europeans (e.g.
harvesting, introduction of game fish and their diseases, habitat modification) likely
led to the only known extinction of an endemic NZ freshwater fish, the NZ grayling
(Prototroctes oxyrhynchus). The museum trade potentially also contributed to
further extinctions (e.g. Auckland Island merganser Mergus australis), even while
species protections were being put in place (Worthy and Holdaway 2002).

The faunal extinctions created vacancies in ecological niches. In some cases,


ancient DNA and radiocarbon-dating has shown these niches were subsequently
filled in biological turnover events by closely related taxa or a different genetic
lineage of the original taxon. To date, these previously cryptic events have been
observed in Megadyptes penguins (Waitaha to yellow-eyed M. antipodes), little
penguins (Eudyptula minor to E. novaehollandiae), sea lions (prehistoric NZ to
subantarctic lineage) and waterfowl (Poūwa Cygnus sumnerensis to black swan
C. atratus; Rawlence et al. 2017; Waters et al. 2017). While the extinctions were
contemporaneous with the first few centuries after human arrival, replacements in
most cases were temporally offset. In contrast, Megadayptes penguins and sea lions
exhibited synchronous biological turnover, with replacement potentially facilitated
by cooler climate and reduced anthropogenic pressure during the Little Ice Age
(Waters et al. 2017). Closely related, open-habitat adapted Australian avifauna also
replaced forest-adapted NZ avifauna as a result of the extinction of NZ congeners
and environmental modification (Rawlence et al. 2019).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
242 Historical Ecology

While numerous NZ species went extinct, many others were reduced to relict
populations, coincident in some cases with significant reductions in genetic and
morphological diversity (e.g. Scarsbrook et al. 2022). One such species, the South
Island takahē (Porphyrio hochstetteri) was considered extinct until 1948 CE when it
was rediscovered. Much of NZ’s herpetofauna, especially larger species (e.g.
Hamilton’s frog Leiopelma hamiltoni, tuatara Sphenodon punctatus and Duvaucel’s
gecko Hoplodactylus duvaucelii) found refuge on offshore islands following
Polynesian arrival and mainland extirpation (Worthy and Holdaway 2002). Ancient
DNA research is highlighting the impacts of these bottlenecks. kākāpō (Strigops
habroptilus) underwent two bottlenecks associated with Polynesian, and later,
European arrival (Seersholm et al. 2018). Otago shag (Leucocarbo chalconotus) lost
99% of its genetic variation within a century of Polynesian arrival (Rawlence et al.
2016b), while fur seals exhibited significant changes in haplotype frequency
following contraction into and expansion out of refugia (Salis et al. 2016). However,
not all range-contractions were genetically detrimental – Fiordland crested penguin
(Eudyptes pachyrhynchus) retained much of its genetic diversity, despite retraction
into refugia (Cole et al. 2019).

Human arrival resulted in not just extinctions, but also lost ecological
connections. NZ’s flora exhibits a high proportion of hypothesized anti-browsing
traits (e.g. divarication, heteroblasty, spines, unpalatable leaves and toxins). Plant
species with these traits have been found in moa coprolites, lending support to the
anti-browsing hypothesis and loss of selective pressure post extinction (Wood et al.
2008). NZ’s avifauna may also have been important pollinators and seed-dispersers.
Ancient DNA and palynological analysis of kākāpō coprolites showed a high
abundance of pollen from the parasitic wood rose (Dactylanthus taylorii),
suggesting that kākāpō (now restricted to offshore islands) may have fed on and
pollinated this endangered plant when compared with previously only known native
pollinator, the lesser short-tailed bat (M. tuberculata; Wood et al. 2012). kākāpō and
moa also consumed colorful fruiting ectomycorrhizal fungi and may have played a
role in fungal dispersal (Boast et al. 2018). Paleodietary evidence also suggests moa
were important small seed (<3.3 mm) dispersers of herbs and shrubs – large seeds
do not survive passage through the gizzard (Wood et al. 2008). The extinctions and
population bottlenecks experienced by NZ’s avifauna may have also resulted in the
co-extinction of associated parasite faunas, although further research is needed to
determine if parasites found in moa and kākāpō coprolites survive in other avian
herbivores (Boast et al. 2018).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reconstructing the Impact of Humans on Aotearoa New Zealand’s Biodiversity 243

18.6. Conclusion

The arrival of Polynesians, and later Europeans, in Aotearoa, set in motion a


series of events that would fundamentally change the NZ ecosystem.
Multidisciplinary research has been at the forefront of shedding light on this
dynamic period of NZ’s history, and is answering some of the most pertinent
questions in archaeology and paleoecology. However, many questions remain
unanswered, and await further research and new analytical techniques. As in
geology, the past is the key to the present, and we would argue, also the future.

18.7. References

Allentoft, M.E., Heller, R., Oskam, C.L., Lorenzen, E.D., Hale, M.L., Gilbert, M.T.P.,
Jacomb, C., Holdaway, R.N., Bunce, M. (2014). Extinct New Zealand megafauna were
not in decline before human colonization. PNAS, 111, 4922–4927.
Anderson, A. (1989). Prodigious Birds: Moas and Moa-Hunting in Prehistoric New Zealand.
Cambridge University Press, Cambridge.
Armstrong, T.J. and de Lange, P.J. (2005). Conservation genetics of Hebe speciosa
(Plantaginaceae) an endangered New Zealand shrub. Bot. J. Linn. Soc., 149, 229–239.
Atherton, R.A., Lockhart, P.J., McLenachan, P.A., de Lange, P.J., Wagstaff, S.J.,
Shepherd, L.D. (2015). A molecular investigation into the origin and human-mediated
spread of karaka/kōpi (Corynocarpus laevigatus) in New Zealand. J. Roy. Soc. NZ., 45,
212–220.
Boast, A.P., Weyrich, L.S., Wood, J.R., Metcalf, J.L., Knight, R., Cooper, A. (2018).
Coprolites reveal ecological interactions lost with the extinction of New Zealand
birds. PNAS, 115, 1546–1551.
Cole, T.L., Rawlence, N.J., Dussex, N., Ellenberg, U., Houston, D.M., Mattern, T.,
Miskelly, C.M., Morrison, K.W., Scofield, R.P., Tennyson, A.J.D. et al. (2019). Ancient
DNA of crested penguins: Testing for temporal genetic shifts in the world’s most diverse
penguin clade. Mol. Phylo. Evol., 131, 72–79.
Diamond, J.M. (1990). New Zealand as an archipelago: An international perspective. Ecol.
Restor. NZ. Is., 2, 3–8.
Duncan, R.P. and Blackburn, T.M. (2004). Extinction and endemism in the New Zealand
avifauna. Global Ecol. Biogeog., 13, 509–517.
Greig, K. and Rawlence, N.J. (2021). The contribution of kurī (Polynesian dog) to the
ecological impacts of the human settlement of Aotearoa New Zealand. Front. Ecol. Evol.,
9, 757988.
King, M. (2003). Penguin History of New Zealand. Penguin, Rosedale.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
244 Historical Ecology

Maxwell, J.J. and Tromp, M. (2016). Corynocarpus laevigatus: Where art thou? Finding
evidence of this elusive tree crop. Rev. Palaeobot. Palynol., 234, 198–210.
Maxwell, J.J., Howarth, J.D., Vandergoes, M.J., Jacobsen, G.E., Barber, I.G. (2016). The
timing and importance of arboriculture and agroforestry in a temperate East Polynesia
society, the Moriori, Rekohu (Chatham Island). Quat. Sci. Rev., 149, 306–325.
McDowall, R.M. (2011). Ikawai: Freshwater Fishes in Māori Culture and Economy.
Canterbury University Press, Christchurch.
McGlone, M.S., Wilmshurst, J.M., Leach, H.M. (2005). An ecological and historical review
of bracken (Pteridium esculentum) in New Zealand, and its cultural significance. NZ. J.
Ecol., 29, 165–184.
Perry, G.L.W., Wheeler, A.B., Wood, J.R., Wilmshurst, J.M. (2014a). A high-precision
chronology for the rapid extinction of New Zealand moa (Aves, Dinornithiformes). Quat.
Sci. Rev., 105, 126–135.
Perry, G.L.W., Wilmshurst, J.M., McGlone, M.S. (2014b). Ecology and long-term history of
fire in New Zealand. NZ. J. Ecol., 38, 157–176.
Powell, A.W.B. (1938). The Paraphantidae of New Zealand, No. IV and the genus
Placostylus in NZ. Rec. Auck. Inst. Mus., 2, 141–150.
Rawlence, N.J., Collins, C.J., Anderson, C.N.K., Maxwell, J.J., Smith, I.W.G.,
Robertson, B.C., Knapp, M., Horsburgh, K.A., Stanton, J.A.L., Scofield, R.P. et al.
(2016a). Human-mediated extirpation of the unique Chatham Islands sea lion and
implications for the conservation management of remaining New Zealand sea lion
populations. Mol. Ecol., 25, 3950–3961.
Rawlence, N.J., Scofield, R.P., Spencer, H.G., Lalas, C., Easton, L.J., Tennyson, A.J.D.,
Adams, M., Pasquet, E., Fraser, C., Waters, J.M. et al. (2016b). Genetic and
morphological evidence for two species of Leucocarbo shag (Aves, Pelecaniformes,
Phalacrocoracidae) from southern South Island of New Zealand. Zool. J. Linn. Soc., 177,
676–694.
Rawlence, N.J., Kardamaki, A., Easton, L.J., Tennyson, A.J.D., Scofield, R.P., Waters, J.M.
(2017). Ancient DNA and morphometric analysis reveal extinction and replacement of
New Zealand’s unique black swans. Proc. B., 284, 20170876.
Rawlence, N.J., Scofield, R.P., McGlone, M.S., Knapp, M. (2019). History repeats: Large
scale synchronous biological turnover in avifauna from the Plio-Pleistocene and Late
Holocene of New Zealand. Front. Ecol. Evol., 7, 158.
Ross, P.M., Know, M.A., Smith, S., Smith, H., Williams, J., Hogg, I.D. (2018). Historical
translocations by Māori may explain the distribution and genetic structure of a threatened
surf clam in Aotearoa (New Zealand). Sci. Rep., 8, 17241.
Salis, A.T., Easton, L.J., Robertson, B.C., Gemmell, N., Smith, I.W.G., Weisler, M.I.,
Waters, J.M., Rawlence, N.J. (2016). Myth or relict: Does ancient DNA detect the
enigmatic Upland seal? Mol. Phylo. Evol., 97, 101–106.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reconstructing the Impact of Humans on Aotearoa New Zealand’s Biodiversity 245

Scarsbrook, L., Verry, A.J.F., Walton, K., Hitchmough, R.A., Rawlence, N.J. (2022). Ancient
mitochondrial genomes recovered from small vertebrate bones through minimally
destructive DNA extraction: Phylogeography of the New Zealand gecko genus
Hoplodactylus. Mol. Ecol., doi.org/10.1111/mec.16434 [in press].
Seersholm, F.V., Cole, T.L., Grealy A., Rawlence, N.J., Greig, K., Knapp, M., Stat, M.,
Hansen, A.J., Easton, L.J., Shepherd, L. et al. (2018). Subsistence practices, past
biodiversity, and anthropogenic impacts revealed by New Zealand-wide ancient DNA
survey. PNAS, 115, 7771–7776.
Shepherd, L.D., Bulgarella, M., de Lange, P.J. (2018). Genetic structuring of the coastal herb
Arthropodium cirratum (Asparagaceae) is shaped by low gene flow, hybridization and
prehistoric translocation. PLoS ONE, 13, e0204943.
Smith, I.W.G. (1989). Maori impact on the marine megafauna: Pre-European distributions of
New Zealand sea mammals. In Saying So Doesn’t Make It So: Papers in Honour of B.
Foss Leach, Sutton, D.G. (ed.). New Zealand Archeological Association, Dunedin.
Steadman, D.W. (2006). Extinction and Biogeography of Tropical Pacific Birds. University
of Chicago Press, Chicago.
Walter, R., Buckley, H., Jacomb, C., Matisoo-Smith, E.A. (2017). Mass migration and the
Polynesian settlement of New Zealand. J. World Prehist., 30, 351–376.
Waters, J.M., Fraser, C.I., Maxwell, J.J., Rawlence, N.J. (2017). Did interaction between
human pressure and Little Ice Age drive biological turnover in New Zealand?
J. Biogeog., 44, 1481–1490.
Wehi, P.M. and Clarkson, B.D. (2007). Flora of New Zealand 10. Phormium tenax, harakeke,
New Zealand flax. NZ. J. Bot., 45, 521–544.
Wilmshurst, J.M., Anderson, A.J., Higham, T.F.G., Worthy, T.H. (2008). Dating the late
prehistoric dispersal of Polynesians to New Zealand using the commensal Pacific rat.
PNAS, 105, 7676–7680.
Wood, J.R., Rawlence, N.J., Rogers, G.M., Austin, J.J., Worthy, T.H., Cooper, A. (2008).
Coprolite deposits reveal the diet and ecology of the extinct New Zealand megaherbivore
moa (Aves, Dinornithiformes). Quat. Sci. Rev., 27, 2593–2602.
Wood, J.R., Wilmshurst, J.M., Worthy, T.H., Holzapfel, A.S., Cooper, A. (2012). A lost link
between a flightless parrot and a parasitic plant and the potential role of coprolites in
conservation paleobiology. Conserv. Biol., 26, 1091–1099.
Worthy, T.H. and Holdaway, R.N. (2002). The Lost World of the Moa: Prehistoric Life of
New Zealand. Canterbury University Press, Christchurch.
Worthy, T.H., De Pietri, V., Scofield, R.P. (2017). Recent advances in avian palaeobiology in
New Zealand with implications for understanding New Zealand’s geological, climatic and
evolutionary histories. NZ. J. Zool., 44, 177–211.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
19

Historical Ecology of the


Coastal Aeolian Sedimentary
Systems of the Canary Islands
Aarón Moisés SANTANA-CORDERO, Antonio Ignacio HERNÁNDEZ-
CORDERO, Néstor MARRERO-RODRÍGUEZ, Leví GARCÍA-ROMERO,
Elisabet FERNÁNDEZ-CABRERA, Carolina PEÑA-ALONSO, Emma
PÉREZ-CHACÓN ESPINO and Luis HERNÁNDEZ-CALVENTO
Universidad de Las Palmas de Gran Canaria, Canary Islands, Spain

19.1. Introduction

A brief history is presented in this work of the evolution of the coastal aeolian
sedimentary systems of five islands of the Canary Archipelago in Spain (Tenerife,
Gran Canaria, Fuerteventura, Lanzarote and La Graciosa). Taking the Spanish
conquest of the islands as the starting point, the aim is to provide a synthesis of the
500-year-old relationship between Canary society and some of the island’s aeolian
ecosystems. This will help to explain and understand their present-day status and
functioning. An analysis is undertaken of the interactions between the natural
dynamics of these systems and the land-uses that have been developed in them.
Historical sources were used for the period which extends from the 1500s to the
1900s, mainly archived documents, cartography and secondary sources which are

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
248 Historical Ecology

reflected in the references. For more recent times (20th and 21st centuries),
additional information was obtained from field work, historical aerial photographs
and recent digital orthophotos. In both cases, the spatio-temporal analyses were
principally carried out through the use of geographic information systems (GIS) to
determine the environmental changes on the basis of biotic and geomorphological
variables.

The role of these environments in the socioeconomic development of the Canary


Islands has been crucial and it helps explaining the growing interest in recent
decades in the coastal ecosystems of the Canary Islands (Spain). These systems
constitute an ideal laboratory for the study of the consequences of the interactions
between their natural dynamics and land-uses, particularly in reference to arid island
environments. The aboriginal inhabitants of the islands had lived a pre-European
existence for only around 1,500 years before they were conquered over the course of
the 15th century by the Spanish. The relatively short period of human intervention
since then is enriched by multiple information sources on natural features and
land-uses. In short, the spatio-temporal limits are well known and the information
sources varied.

19.2. Study sites

The coastal aeolian sedimentary systems occupy a total surface area of 18,810 ha
(2.5% of the total land area of the archipelago) and are situated mainly on the
eastern islands (Hernández-Cordero et al. 2019). The fact that these systems
developed on oceanic islands, in an arid climate, and have been subjected in recent
decades to intensive human activity in the form of year-round tourism is
distinguishing eco-anthropic features in a European context. The climate conditions
are characterized by annual precipitation of just 81–110 mm, mean annual
temperatures ranging between 19.7 and 21°C, and the important influence of the
trade winds which blow from the NE. The desiccating effect of the wind and the low
amount of rainfall have resulted in low density and mainly shrubby vegetation
(Hernández-Cordero et al. 2019) which, in turn, has a considerable impact on the
geomorphological processes. Consequently, these systems are transgressive and with
the potential to be highly dynamic. The aeolian landforms, conditioned by the
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Ecology of the Coastal Aeolian Sedimentary Systems of the Canary Islands 249

vegetation and sediment supply, range from nebkhas to transverse ridges. Their high
complexity, fragile nature, uniqueness and importance from the point of view of
biodiversity and geodiversity explain why most of them have been granted
protection status through different local, regional, national and international (EU)
regulations.

These systems have evolved differently depending on their individual natural


characteristics and the degree and type of anthropic pressure to which they have
been subjected. Below, a joint analysis is presented of these aeolian sedimentary
systems from the perspective of historical ecology. The systems considered in this
study are the following: the coastal aeolian sedimentary systems of La Graciosa
Island, El Jable (Lanzarote), El Médano (Tenerife), Jandía and Corralejo
(Fuerteventura), and Maspalomas and Guanarteme (Gran Canaria) (Figures 19.1 and
19.2).

Figure 19.1. Some of the sedimentary systems studied. Top-left: Jandía, top-right:
Corralejo (by Claudio Moreno Medina), bottom-left: El Médano and bottom-right:
Maspalomas. For a color version of this figure, see www.iste.co.uk/decocq/ecology.
zip
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
250 Historical Ecology

Figure 19.2. Study sites. Old and current status of the arid coastal aeolian
sedimentary systems studied. The time intervals considered correspond to those
used in the studies conducted in García-Romero et al. (2016), Santana-Cordero et al.
(2016a, 2016b) and Marrero-Rodríguez et al. (2020a, 2020b). For a color version of
this figure, see www.iste.co.uk/decocq/ecology.zip
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Ecology of the Coastal Aeolian Sedimentary Systems of the Canary Islands 251

19.3. Historical evolution of the coastal aeolian sedimentary systems of


the Canary Islands

The anthropic impact on the aeolian sedimentary systems is directly related to


the productive models that have been in place over the course of the islands’ history
since their conquest by the Spanish. This will therefore form the basis of the line of
reasoning that will be followed in the presentation of the transformations that these
systems have undergone due to changes in land-use.

In 1730–1736, the volcanic eruption of Timanfaya (Lanzarote) buried a


significant amount of land used for agrarian purposes, including crop cultivation,
grazing and firewood gathering. One result was the intensification of land-uses in
El Jable (Lanzarote) and La Graciosa, while another was an increase in migration to
Corralejo (Fuerteventura). After this eruption, herd grazing on La Graciosa became a
permanent activity and was accompanied by the first attempts at the establishment of a
population settlement. The consequent pressure that was imposed on the environment
led to the issuing of local ordinances and regulations which, among other things,
prohibited permanent settlements, the cutting or burning of shrubs, and the
unauthorized entrance of livestock. A similar deforestation process was taking place in
El Jable (Lanzarote) related, in this case, to the burning of plant material to
manufacture liquors and the plowing of land for crop growing (Caballero-Mújica
1991). As a result of these processes, sediment remobilization took place through the
action of aeolian sedimentary dynamics. In El Jable (Lanzarote), for example, there are
descriptions of large mounds of sand crossing the island from north to south (Torriani
1959), while in Guanarteme (Gran Canaria), the historical cartography evidences the
presence of landforms from 1599 and, over a century later, the change in orientation
between 1742 and 1792 of a large dune in its northern sector (Santana-Cordero et al.
2014). In other systems, however, similar processes do not appear to have taken
place, certainly not at any significant scale. For example, in Maspalomas (Gran
Canaria), with some 100 inhabitants towards the end of the 18th century (Marrero
and Naranjo 1999), the few land-uses that existed were non-intensive, and no
significant environmental transformations seem to have been generated.

In general, the studied aeolian sedimentary systems were perceived socially from
two distinct perspectives, which caused different anthropogenic pressures on them.
On the one hand, some of the systems were seen as marginal territories due to their
distance from the main population settlements and their aridity (lack of water
resources) and consequent unsuitability for agriculture, especially in Gran Canaria
and Tenerife (cases of Maspalomas and El Médano) which enjoyed considerably
more abundant agroforestry resources. For example, the location of Maspalomas, at
the southern tip of Gran Canaria and at some distance from the island’s main
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
252 Historical Ecology

population centers, accentuated its isolation and, possibly, the continued


preservation of many of its natural features. On the other hand, in the dune systems
of Lanzarote, Fuerteventura and La Graciosa, the general characteristics of aridity,
low height and scant wood resources provoked the establishment of more intensive
land-uses in their aeolian sedimentary systems (Marrero-Rodríguez et al. 2020a), as
well as the introduction of more profound modifications. Exceptionally, Guanarteme
was seen from both these perceptions during its history, i.e. as a marginal place first,
and subsequently as an essential place for occupation.

19.3.1. 19th century: territorial consolidation and spread of the agrarian


socioeconomic system

Over the course of this period, the evolution of the aeolian sedimentary systems
situated on islands with fewer agrarian resources (and where a relative
overexploitation of these marginal areas took place) differed from the evolution of
other systems located on islands with more resources. Another differentiating aspect
was the development of the limestone industry around some of the systems and the
consequent exploitation of the resources required for its practice. Indeed, this
industry had perhaps one of the biggest impacts on the systems during this period, as
not only did the limestone itself have to be quarried to provide the raw material
(basically calcareous rock from coastal paleo-levels underlying loose aeolian
sediments), but vegetation also had to be cut and gathered from the area to fire the
limekilns. This deforestation led in turn to the remobilization of a significant volume
of sediment and major changes to the landforms of these systems. The impact of this
activity was especially notable in Jandía, Fuerteventura (Marrero-Rodríguez et al.
2020a). Together with extensive livestock grazing (mainly goats and sheep), this
activity severely impacted the vegetation of the area. The effects of goat and sheep
grazing were also significant in the aeolian system of Corralejo, in the north of the
same island.

Another example of relative environmental overexploitation took place in La


Graciosa, where there is evidence of a drastic reduction in plant cover
(Santana-Cordero et al. 2016a) and generalized sediment remobilization processes.
This was due not only to the use of limekilns, but also because of a notable
population increase from 1880 onwards which added pressure on the environment
through grazing, plowing, sand extraction for construction purposes, etc. In the case
of El Jable (Lanzarote), a sand-based crop system was even established for which an
extensive sector of the aeolian sedimentary system was prepared using plant-based
windbreaks to slow down the circulation of sediment.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Ecology of the Coastal Aeolian Sedimentary Systems of the Canary Islands 253

For its part, the aeolian sedimentary system of Guanarteme (Gran Canaria) had
managed to retain many of its natural features despite bordering the island’s capital,
with 8.35–11.14 m high dunes being described in 1815 (von Buch 1999). By the
mid-point of the century, however, several land-uses can be found in this system
(Santana-Cordero et al. 2016b), including for agricultural, residential, recreational
and extraction purposes. The consequences were modifications to a system that was
perceived by the society of the time as an enemy that had to be conquered, as the
sands were invading crops and road networks and approaching the entrance to the
city. Contrastingly, in the southern area of the island, the Maspalomas dune system
remained in a good state of conservation, principally because of the continued lack
of interest in that space and, consequently, the low levels of anthropic pressure on
the dune field (Marrero and Naranjo 1999).

19.3.2. 20th century to the present day: the tourism transformation

The evolution of the different aeolian sedimentary systems considered in the


present study varied substantially between the first half of the 20th century and the
following decades when the urbanization process and the development of tourism
became more widespread (Figure 19.1), albeit at differing rates and intensities.

In the first 50 years of the last century, there was a gradual decline in agrarian
land-use in some of the aeolian sedimentary systems. At the same time, the import of
fossil fuels began to reduce the pressure on vegetation, allowing plant recolonization
and the reappearance of nebkhas in some islands, as occurred in El Jable (Lanzarote)
and La Graciosa. However, in Jandía (Fuerteventura), oral sources suggest the
opposite process was taking place in the period 1940–1960, with intensive
deforestation taking place (Marrero-Rodríguez et al. 2020a). This can be attributed
to a simultaneous increase in grazing and in the demand for firewood for the
limestone industry. The aeolian sedimentary system was significantly destabilized
and plant regeneration so affected that limestone began to be exported as raw
material because there was insufficient plant material to fire the limekilns. At around
the same time, intensive tomato growing became established between the 1930s and
1960s in Maspalomas (Gran Canaria), which meant the plowing and preparation of
80.9 ha situated in the northern sector of the dune field. This new activity
additionally resulted in a population increase as people began to migrate from the
north to the south of the island in search of work, with the number of inhabitants
rising to 13,384.

However, the factor that by some distance most impacted the socioeconomic
system of the islands was the emergence of tourism. A radical change in land-use
took place from the 1960s onwards in Maspalomas (Gran Canaria), as well as a
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
254 Historical Ecology

significant revaluation on the part of society of two resources which were key to the
emergence of mass tourism, sun and beach, which had until then been virtually
ignored. The development of the area as a tourist resort continued throughout the
1970s, 1980s and 1990s (Parreño Castellano 2001), with the area surrounding the
dune field occupied by buildings and infrastructures of various kinds, as well as a
golf course. This resulted in a major alteration to the aeolian sedimentary dynamics
and, consequently, various ecological processes (Hernández-Calvento et al. 2014;
Hernández-Cordero et al. 2017, 2018). In Corralejo (Fuerteventura), although the
first bungalows were constructed in 1962, it was not until 1975 and the construction
of two large hotels inside the aeolian sedimentary system that the tourist industry
began to make its presence felt on the environment. Likewise, in El Jable
(Lanzarote), a significant change in the economic model emerged, with traditional
activities giving way to tourism and provoking a dramatic increase in anthropic
pressure on the dune system. The first tourists began to arrive in El Médano
(Tenerife) in 1964, and it was not long before cultivation of the tomato was
gradually abandoned. Finally, La Graciosa also initiated an orientation of its
economy towards mass tourism, with an almost doubling of the built-up area for its
population center between 1977 and 2009 and a considerable extension of the
island’s port area.

Tourism entailed enormous changes to the economic and social structure, as well
as to various landscape features (most notably in Maspalomas, Corralejo, Jandía, El
Médano and La Graciosa). The construction of housing, tourist complexes and all
manner of infrastructure resulted in the transformation of these systems and
important environmental consequences (García-Romero et al. 2016; Table 19.1):
(i) a decrease in the surface area of aeolian sedimentary systems (surface variation %),
particularly in Maspalomas and El Médano, (ii) an increase in plant cover
(vegetation variation %) and (iii) a decrease in the surface area occupied by bare
sand (bare sand variation %).

In addition, a reduction of the distribution in the systems of some plant


communities took place. For example, in Maspalomas, studies have been conducted
on the reduction of the area covered by Traganum moquinii (-24.4%), essential for
foredune formation (García-Romero et al. 2021), and Plocama pendula (-99.8%), as
well as on the total disappearance of the Euphorbia balsamifera community
(Hernández-Cordero et al. 2017). In other systems, such as Guanarteme, the total
disappearance of T. moquinii communities has been confirmed (Santana-Cordero
et al. 2014), while in La Graciosa, this plant community was severely impacted by
its use as firewood for cooking (Santana-Cordero et al. 2016a). A dramatic reduction
took place in Jandía in the number of specimens of Convolvulus caput-medusae
because of its use as a fuel (Marrero-Rodríguez et al. 2020a), while the reduction of
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Ecology of the Coastal Aeolian Sedimentary Systems of the Canary Islands 255

T. moquinii in El Médano was due to the opening of extractive quarries which


generated flooded areas where the plant was unable to grow (Marrero-Rodríguez
et al. 2020b). An important coastal erosion also occurred in Maspalomas and Jandía,
evidenced in the regression of the coastline (Alonso et al. 2002; Hernández-Cordero
et al. 2018). Meanwhile, the Guanarteme system had been completely buried by the
1960s as a result of the incessant expansion of the city of Las Palmas de Gran
Canaria (Santana-Cordero et al. 2014).

Surface Bare sand


Vegetation
System/indicators variation variation
variation (%)
(%) (%)
Maspalomas (Gran Canaria) (1961–2003) -24.0 38.8 -61.3
Corralejo (Fuerteventura) (1969–2003) -12.9 8.9 -32.0
Lambra (La Graciosa) (1954–2009) 0.0 12.1 -47.1
Jable Sur (La Graciosa) (1954–2009) -1.5 60.9 -77.5
Jandía (Fuerteventura) (1963–2016) -3.6 50.1 -62.9
El Médano (Tenerife) (1964–2018) -50.9 -18.5 -6.2

Table 19.1. Main land cover changes over the past 70 years in the aeolian
sedimentary systems of the Canary Islands. Modified from García-Romero et al.
(2016) and Marrero-Rodríguez et al. (2020a, 2020b)

Eventually, a growing awareness of the pressure being exerted on these systems


through urban and tourism developments, as well as of the value and importance of
the biodiversity and geodiversity of these ecosystems, led to them being declared
protected nature reserves (Act 12/1987, dated June 19, on Natural Protected Areas in
the Canary Islands), with the exception of a major part of the El Jable system and
the Guanarteme system which had already been built over. This legislation aimed
not only to preserve what remained of these ecosystems by preventing further
construction in them but also provided regulations on traditional uses, giving rise to
the natural regeneration of the vegetation and, to a certain degree, of the ecological
processes in most of the systems (Table 19.1).

19.4. Conclusion

Over the course of the historical period considered in this study, different
land-uses associated with the prevailing economic systems of the time have been
developed in the aeolian sedimentary systems of the Canary Islands. The transition
from an agrarian-based to a tourism-based system resulted in the decline of some
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
256 Historical Ecology

uses and the emergence of other new ones. The traditional activities of an agrarian
society dominated between the 15th century and the start of the 20th century.
However, the uses that were made of most of the dune systems were generally
marginal, except on the islands of Lanzarote and Fuerteventura given the limited
resources that these systems contained for an agrarian society. In this regard, the
case of Guanarteme was unique, impacted at an earlier stage by the growth and
expansion of the city of Las Palmas de Gran Canaria.

Land-use changed radically in the second half of the 20th century as a result of
urban development and the rise of tourism, with significant environmental changes
taking place: (i) the disappearance of the Guanarteme dune system, (ii) a reduction
in surface area in some of the aeolian sedimentary systems and (iii) ecological and
geomorphological change, related to a decrease in the numbers of some plant
communities, increased plant cover, sediment stabilization and the emergence of
coastal erosion processes.

In summary, these ecosystems are no longer considered an “inconvenience” and


of little interest for an agrarian society. Instead, they have become a fundamental
resource for the growth in tourism and the development of the Canary economy over
the course of the last 70 years.

19.5. References

Alonso, I., Alcántara-Carrió, J., Cabrera, L. (2002). Tourist resorts and their impact on beach
erosion at Sotavento beaches, Fuerteventura, Spain. J. Coastal Res., SI 36, 1–7.
von Buch, L. (1999). Descripción física de las Islas Canarias. Graficolor, La Laguna.
Caballero-Mújica, F. (ed.) (1991). Compendio brebe y fasmosso, histórico y político, en que
(se) contiene la cituazion, población, división, gobierno, produziones, fábricas y comercio
que tiene la ysla de Lanzarote en el año de 1776. Ayuntamiento de Teguise, Teguise-Las
Palmas.
García-Romero, L., Hernández-Cordero, A.I., Fernández-Cabrera, E., Peña-Alonso, C.,
Hernández-Calvento, L., Pérez-Chacón, E. (2016). Urban-touristic impacts on the aeolian
sedimentary systems of the Canary Islands: Conflict between development and
conservation. Isl. Stud. J., 11(1), 91–112.
García-Romero, L., Hernández-Cordero, A.I., Hesp, P.A., Hernández-Calvento, L., Santana
del Pino, A. (2021). Decadal monitoring of Traganum moquinii’s role on foredune
morphology of an human impacted arid dunefield. Sci. Total Environ., 758, 143802.
Hernández-Calvento, L., Jackson, D.W.T., Medina, R., Hernández-Cordero, A.I., Cruz, N.,
Requejo, S. (2014). Downwind effects on an arid dunefield from an evolving urbanised
area. Aeolian Res., 15, 301–309.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Ecology of the Coastal Aeolian Sedimentary Systems of the Canary Islands 257

Hernández-Cordero, A.I., Hernández-Calvento, L., Pérez-Chacón Espino, E. (2017).


Vegetation changes as an indicator of impact from tourist development in an arid
transgressive coastal dune field. Land Use Policy, 64, 479–491.
Hernández-Cordero, A.I., Hernández-Calvento, L., Hesp, P.A., Pérez-Chacón, E. (2018).
Geomorphological changes in an arid transgressive coastal dune field due to natural
processes and human impacts. Earth Surf. Proc. Land., 43(10), 2167–2180.
Hernández-Cordero, A.I., Peña-Alonso, C., Hernández-Calvento, L., Ferrer-Valero, N.,
Santana-Cordero, A.M., García-Romero, L., Pérez-Chacón Espino, E. (2019). Aeolian
sedimentary systems of the Canary Islands. In The Spanish Coastal Systems, Morales,
J.A. (ed.). Springer, Cham.
Marrero, J.L. and Naranjo, R. (1999). Maspalomas espacio natural. Ayuntamiento de San
Bartolomé de Tirajana (Concejalía de Turismo), Las Palmas de Gran Canaria.
Marrero-Rodríguez, N., García-Romero, L., Sánchez-García, M.J., Hernández-Calvento, L.,
Pérez-Chacon Espino, E. (2020a). An historical ecological assessment of land-use
evolution and observed landscape change in an arid aeolian sedimentary system. Sci.
Total Environ., 716, 137087.
Marrero-Rodríguez, N., García-Romero, L., Peña-Alonso, C., Hernández-Cordero, A.
(2020b). Biogeomorphological responses of nebkhas to historical long-term land uses in
an arid coastal aeolian sedimentary system. Geomorphology, 368, 107348.
Parreño Castellano, J.M. (2001). El proceso de urbanización del espacio turístico. In
Evolución e implicaciones del turismo en Maspalomas Costa Canaria (tomo I), vol. 1,
Hernández Luis, J.A. and Parreño Castellano, J.M. (eds). Ayuntamiento de San Bartolomé
de Tirajana (Concejalía de Turismo), San Bartolomé de Tirajana.
Santana-Cordero, A.M., Monteiro-Quintana, M.L., Hernández-Calvento, L. (2014).
Reconstructing the environmental conditions of extinct coastal dune systems using
historical sources: The case of the Guanarteme dune field (Canary Islands, Spain). J.
Coast. Conserv., 18, 323–337.
Santana-Cordero, A.M., Monteiro Quintana, M.L., Hernández Calvento, L., Pérez-Chacón
Espino, E., García-Romero, L. (2016a). Long-term human impacts on the coast of La
Graciosa, Canary Islands. Land Degrad. Dev., 27, 479–489.
Santana-Cordero, A.M., Monteiro-Quintana, M.L., Hernández-Calvento, L. (2016b).
Reconstruction of the land uses that led to the termination of an arid coastal dune system:
The case of the Guanarteme dune system (Canary Islands, Spain), 1834–2012. Land Use
Policy, 55, 73–85.
Torriani, L. (1959). Descripción e historia del reino de las Islas Canarias antes Afortunadas,
con el parecer de sus fortificaciones (Vol. 2). Goya Ediciones, Santa Cruz de Tenerife.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
20

Historical Forest Microclimates


Emiel DE LOMBAERDE1, Karen DE PAUW1, Pallieter DE SMEDT1,
Jonathan LENOIR2, Camille MEEUSSEN1, Thomas VANNESTE1,
Kris VERHEYEN1, Florian ZELLWEGER3 and Pieter DE FRENNE1
1
Ghent University, Belgium
2
Jules Verne University of Picardie, Amiens, France
3
Swiss Federal Institute for Forest, Snow and Landscape Research, Birmensdorf, Switzerland

The majority of terrestrial organisms live in forests where microclimatic


conditions below trees differ significantly from conditions outside forests. The
temperature, humidity, wind and light environment in forest understories are
strongly affected by the structure and composition of the forest, local topography,
landscape characteristics and the regional macroclimatic conditions, among others,
leading to highly variable microclimates in structurally contrasting forests and along
successional or management gradients (Geiger et al. 2003). Temperature extremes
are often strongly buffered in forests compared to open habitats, with cooler below –
canopy maximum temperatures, warmer minimum temperatures and lower seasonal
and interannual variability (De Frenne et al. 2021).

Climate warming, land-use change and the history of forest management are
important determinants of the temporal dynamic of forest microclimates. However,
very little is known about the history (and future) of forest microclimates, i.e. how
microclimate conditions in the understory of forests are changing over time. The
knowledge gap on the history of forest microclimates chiefly stems from a lack of
long-term forest microclimate time series. Long-term climatic time series are almost
exclusively based on weather stations located outside forests in open landscapes
(with some exceptions such as the FLUXNET, ICP forests or rare regional
networks). Given the lack of long-term time series of forest microclimates, current

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
260 Historical Ecology

assessments of climate-change impacts on the biodiversity living in the shade of


trees are thus subject to considerable uncertainty. Reconstructing the historical
microclimatic conditions in forest understories is important to capture the realized
climatic niche of forest species and to improve our understanding of the actual
response of forest biodiversity to direct human pressures on forest ecosystems (e.g.
disturbances due to forest management) as well as indirect human pressures through
anthropogenic climate change. Time series of forest microclimate data are therefore
key for assessing time-lag dynamics in forest communities, as well as for studying
the effect of climate microrefugia on community responses to climate change.

In this chapter, we focus on historical microclimates and how they can help us to
predict the future. We summarize the drivers and effects of past, present and future
climate, land-use and forest management on temporal dynamics in understory
microclimate (summarized in Table 20.1), methods to infer historical microclimates
and outline the implications for forest biodiversity. The focus is on temperatures near
the forest floor (0–1 m), because these temperatures are of high biological importance
for understory plants and animals. The focus on temperature also provides a direct link
to climate warming, although we clearly acknowledge that other microclimatic
parameters, such as soil moisture and light availability, are also important.

Drivers of Spatial Importance Changing Methods to infer the


microclimate scale for forest over past
temperatures microclimates ecological
time?
Tree species Plot +++ Yes Legacy vegetation
surveys, remote
sensing,
dendrochronology
Litter quantity and Plot + Yes Legacy vegetation and
quality soil surveys
Forest density, Plot +++ Yes Legacy vegetation
structure surveys, remote
sensing,
dendrochronology
Tree height Plot + Yes Legacy vegetation
surveys, remote
sensing
Understory Plot + Yes Legacy vegetation
community surveys, remote
composition sensing
Edge effects Patch +++ Yes Historical maps
Slope, aspect, Patch (and +++ No NA
topography landscape)
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Forest Microclimates 261

Drivers of Spatial Importance Changing Methods to infer the


microclimate scale for forest over past
temperatures microclimates ecological
time?
Stand size and Patch +++ Yes Historical maps,
diversity, remote sensing,
management archival sources
system
Soil type, parent Landscape + No NA
material (and patch)
Landscape Landscape ++ Yes Historical maps,
composition remote sensing
Fragmentation per Landscape ++ Yes Historical maps,
se vs. habitat area remote sensing
Macroclimate Landscape +++ Yes Weather stations and
downscaling,
mechanistic models
Water bodies, Landscape ++ No NA
rivers, lakes, sea

Table 20.1. Overview of the drivers of forest microclimate from the plot to the
landscape scale, their importance in affecting microclimate temperatures and
whether the effect of the driving factor can change over ecological time

20.1. Drivers of microclimate at the plot, forest and landscape scale

At a very local scale within a forest stand, the microclimate is modified by a


complex interplay of different local factors: forest structure, tree species composition
and height, the abundance of understory plants and topography. At this scale, the
magnitude of temperature offsets (i.e. the horizontal difference between open and
forest conditions) depends on episodic (e.g. due to silvicultural practices) or cyclic
(e.g. due to the seasonal phenology) changes in forest canopy density and composition,
tree height and understory plants. Tree structure and composition affect wind speeds
and patterns, incoming light quantity and quality, moisture availability and finally also
sub-canopy and soil temperatures (Geiger et al. 2003). Even historical disturbances,
for example, due to temporary canopy openings, are known to affect the understory
communities and cause a shift towards warm-adapted species in forest-floor plant
communities. Tree species composition can directly, and indirectly via the forest
structure, affect understory microclimates. Species differ in tree and crown geometry
and leaf area index, and therefore also in their ability to cast shade. Hence, deep shade
casting trees such as beech (Fagus sylvatica) have a large crown depth, and
consequently low understory light levels. Differences also occur between broadleaved
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
262 Historical Ecology

and coniferous forests, with broadleaved forests showing higher evaporative cooling
and a higher ability to buffer maxima temperatures after canopy closure. Historical
forest transformation of broadleaved forest and afforestation with conifers have
therefore had a considerable impact on forest microclimates. Moreover, trees and
shrubs also affect the soil microclimate via their litter quantity and quality. Thick litter
layers, as found under species with a low litter quality (e.g. beech), further reduce soil
temperature variability and buffer maxima and minima temperatures while reducing
soil evaporation. Finally, also the understory community composition might further
influence the forest floor microclimate via similar mechanisms as for the litter layer, in
addition to regulating below-tree canopy microclimates via transpiration.

At the forest stand scale, the management system and edges (see the next
paragraph) are the most important drivers determining spatial and temporal variation
in microclimates (Table 20.1). First, there is a major difference in microclimate
between coppice and coppice-with-standards (CWS) systems, on the one hand, and
high forest-systems, on the other hand. Coppice and CWS systems are characterized
by clear cuts of the coppice layer every 10 to 20 or 30 years, depending on the
species and/or the management target. These frequent stand-level cuttings have a
strong impact on the microclimate and lead to repeated cycles of low microclimatic
buffering being replaced by periods with high buffering. Similar, but more
tempered, patterns develop in CWS systems where a sparse canopy of standard trees
remains in place all the time. Different microclimatic regimes are encountered in
high forest systems (Figure 20.1), due to the substantially longer rotation times in
between regeneration cuts, ranging from multiple decades to centuries, and due to
the generally smaller size of the regeneration cuts. Stand-scale cuts that strongly
disrupt the microclimatic buffering do occur in clear-cut and shelterwood systems.
However, the frequency of these cuts is much lower than in coppice and CWS
systems. The regeneration cuts in strip cutting, group selection and single-tree
selection systems occur on a smaller spatial scale (Figure 20.1), so that the
microclimatic buffering is more or less maintained. Currently, the vast majority of
forests is managed as high forest, but coppice-based systems were in many places
dominant until the middle of the 20th century. This means that forest microclimates
have undergone a very large shift over the course of the last century. More strongly
buffered forest microclimates are now much more prominent than in the past (e.g.
Zellweger et al. 2020). This microclimatic shift is a major driver for the large shifts
in forest biodiversity that is observed in many places throughout Europe (see
section 20.3). Finally, it should be stressed that the effects of the applied
management system extend beyond the stands where the regeneration cuts take
place. Systems where large cuts are applied (notably coppice, CWS, clear-cut and
shelterwood systems) also generate substantial internal forest edges that cause
important microclimatic changes in adjacent stands.
Figure 20.1. Left panel: schematic representation of five common high-forest management systems (A: clear cut,
B: shelterwood system, C: strip cutting system, D: group selection system and E: single-tree selection system). The buffering
capacity of the forest canopy (black arrow) and resulting microclimates will vary between management systems. For instance,
summer temperatures (Tsummer) will be warmer in forest that have undergone large-scale canopy removal such as in clear
cuts (A) compared to forests where canopy disturbances are limited (E); for winter temperatures (Twinter), the opposite will be
the case. Right panel: the spatial and temporal scale of the regeneration cuts applied in the five management systems. Figures
Historical Forest Microclimates

adapted from den Ouden et al. (2010). For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip
263

Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
264 Historical Ecology

Across the world, forests are becoming smaller due to habitat loss and
fragmentation. The high degree of forest fragmentation results in forest edges
increasingly becoming a dominant landscape feature. Worldwide, 20% of forest area
lies within 100 m of a forest edge, but edge effects can penetrate much deeper into
forest patches depending on the forest type. Forest edges have a very contrasting
microclimate compared to forest interiors with, for example, higher air and soil
temperatures, lower soil moisture, higher light availability and higher wind speed
compared to forest interiors. These microclimatic conditions are induced by the
exposed character of forest edges, and also by their altered vegetation structure. As a
result, forest edges harbor distinct communities of plants, vertebrates, arthropods
and microbes. Forest microclimates are increasingly altered through edge effects
since forest fragmentation and hence edge creation is a still ongoing process. Based
on forest edge research, we can assume that microclimate temperatures in forest
patches have increased in the past due to the increasing proportion of edges
compared to interior forest.

The forest microclimate is not only affected by plot and stand characteristics, but
also by the landscape in which it is embedded. Several characteristics of the
surrounding landscape influence the forest microclimate, either in a direct or indirect
way. Topography is one of the most determining factors of a landscape and can have
a strong impact on the microclimate, but does not change strongly over time. The
same applies to the presence and size of large water bodies. However, other
landscape features are of a dynamic nature and these are of prime importance to
estimate how historical (and future) microclimates at the landscape-scale differ from
today (Table 20.1). Microclimatic buffering is related to the macroclimate, a feature
changing on an even larger scale than the landscape scale. Increasing ambient air
temperatures due to macroclimate change might lead to increased buffering of
maximum temperatures in forests. Furthermore, regional hydrological conditions
and varying levels of soil moisture can affect microclimatic buffering through the
evapotranspiration and the transformation of sensible to latent heat. Higher soil
moisture reduces diurnal temperature variation and warming due to incoming
radiation. As a result, also future changes in local water balances are important:
forests might lose their microclimatic buffering capacity if they become
water-limited. In addition to climate change, land-use change can affect forest
microclimates. For example, urbanization and urban centers induce an urban heat
island. Temperatures are typically higher in urbanized areas. Minimum forest
temperatures are warmer in urban forests. The presence of urban areas has changed
strongly over time and is still growing rapidly, as the global distribution of
impervious surface area more than doubled since 1990. Similarly, temporal changes
in the landscape configuration, for example, configuration of forest patches within
the landscape, in terms of spatio-temporal continuity and connectivity, can
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Forest Microclimates 265

potentially affect the history of local climatic conditions. However, knowledge on


the influence of landscape composition and configuration on forest microclimates is
still limited and should be extended, as landscapes have changed due to
anthropogenic and climate-change drivers and will keep on changing in the future.

20.2. Methods to infer microclimate from the past and predict into the
future

To infer microclimates from the past at a plot scale, we need legacy data on
structural variables and tree composition, height and understory plants. These can
chiefly be derived from: (i) historical surveys, (ii) chronosequences, (iii) remote
sensing data (e.g. back to the 1980s using Landsat imagery) and (iv)
dendrochronological data. In that respect, the legacy of botanists who collected
numerous vegetation plots now stored in vegetation databases (Chytrý et al. 2016;
Verheyen et al. 2017) represents a wealth of information and a unique opportunity to
infer past microclimatic conditions in forest understories and thus potentially for
reconstructing long-term time series of forest management and microclimate
(Zellweger et al. 2020). As historical data on forest microclimates are scarce and
difficult to gather, time series of detailed data on local forest structure and
composition collected within national forest inventories as well as other vegetation
resurvey programs provide unique opportunities to shed light on how forest
microclimates have changed over the past century. One promising route is to link the
horizontal (i.e. keeping the height above ground fixed) offset between current
macroclimate and microclimate conditions to current plot or stand-level forest
attributes, such as canopy cover and canopy composition, and apply the resulting
relationships to the same forest attributes measured in past field surveys, thus
reconstructing the forest microclimate of the past (Zellweger et al. 2020). Similarly,
time series describing changes in forest structure and composition provide important
input parameters for modeling microclimates over time based on mechanistic
processes governing the small-scale variation in temperatures (Maclean et al. 2019).
Likewise, chronosequences and “structural sequences” from young forests on
post-agricultural land, older post-agricultural forests and ancient forests with
contrasting tree ages and structures can, via a space-for-time substitution approach,
allow ecologists to infer microclimates from early- versus late-successional stages.
Forest structural attributes can be obtained from baseline and resurveys of
(quasi-)permanent plots and also from repeated remote sensing (see further). A third
approach is to use dated tree-ring data to reconstruct forest structure and
microclimate. Indeed, the reconstruction of past forest structure and management is
possible via analysis of tree rings and growth time series; for instance, approximate
dates of management conversion, tree felling or extreme droughts can be
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
266 Historical Ecology

interpolated from wood cores. Finally, dynamic forest growth models might be
useful in this context too. Such models are often used for forecasting into the future
(e.g. in relation to global change) albeit hindcasting with the goal of inferring past
microclimates should be possible too but has been rarely done.

Historical maps to document past forest cover allow us to go 150–250 years back
in time in many places across Europe and are crucial in estimating past forest
microclimates. Very old maps, such as the French Cassini maps and the Belgian
Ferraris map, both of the 18th century, can be used to date forest patches and thus
distinguish between ancient and more recent forest patches harboring very different
levels of forest biodiversity and ecosystem services and thus potentially different
forest microclimates. Historical microclimates of forest patches can thus be deduced
from old forest maps, by calculating forest edge length. Forest type-specific and
region-specific gradients of microclimatic variables like soil temperature (see, for
example, Arroyo-Rodríguez et al. 2017) and soil moisture can be used to estimate
microclimatic conditions at forest patch level from the past, at least in small forest
fragments with a high edge-to-interior ratio. Unlike forest edges, historical
management systems are less easy to deduce from historical maps. Typically, a
variety of sources have to be used to reconstruct past management. When available,
archival documents about past management, exploitations and wood sales are the
most obvious source of information, and can be used in conjunction with other
sources such as pollen profiles to reconstruct tree species composition (Szabó 2015).
However, written records are often lacking and, in those cases, we have to rely on
field observations, comprising the composition and structure of the tree layer, the
shape of trees and their growth patterns, assessed via dendrochronological methods.
In large forest fragments, species composition and management are likely to fulfill a
larger role in predicting historical forest microclimatic conditions compared to forest
edge. In addition, forest management and tree species composition are main drivers
in edge structural diversity. Microclimatic gradients in forest edges are, however,
also dependent on the surrounding matrix stressing the importance of taking the
landscape level into account when assessing historical microclimates at forest patch
level.

Mapping and monitoring temporal microclimatic variation at the landscape scale


is key to understanding organismal responses to climate variation, which in turn
helps to predict how climate change will impact biodiversity and ecosystems. Yet,
former approaches to map landscape-level microclimates were bound to several
limitations. For instance, networks of microclimatic sensors provide point-based
measurements while weather stations deliver macroclimatic data, but methods to
effectively interpolate or downscale this data have been lacking so far for forests
(Maclean et al. 2019). Several new techniques to map and model microclimates have
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Forest Microclimates 267

emerged over recent decades. Probably one of the most promising techniques is
remote sensing, producing exceptionally detailed and spatially continuous data
layers that can feed into models of horizontal and vertical microclimatic variation of
broad spatio-temporal scales (Zellweger et al. 2019). Remote sensing data can also
be used to infer microclimates at the plot scale (see above).

First, a relatively simple technique to assess microclimates at the landscape level


is the use of land-use maps that offer, sometimes very detailed, snapshots of the
landscape configuration including the position, size and connectivity of forest
patches at the time they were drawn (see above). Digitizing a chronosequence of
historical land-use maps from different points in time may be used to quantify
spatio-temporal changes in a variety of landscape attributes including the amount of
forest cover and forest continuity. Beyond land-use maps, airborne and spaceborne
(e.g. NASA Global Ecosystem Dynamics Investigation; GEDI) light detection and
ranging or “laser scanning” (LiDAR) is now increasingly applied in ecological
studies for a 3D analysis of the Earth’s surface and vegetation (Zellweger et al.
2019). LiDAR data can then be combined with sensor networks providing in situ
measurements of microclimatic conditions on the ground, using statistical
approaches to predict microclimates across the entire LiDAR-mapped area (Frey
et al. 2016a). This technique has been proven effective to model radiation regimes
along vertical canopy profiles in forests as well as light transmittance to the forest
understory at any time in the day or year (Tymen et al. 2017).

Alternatively, aerial photography can be used to map ground topography and


vegetation structure via photogrammetry or structure-for-motion techniques.
Thermal imaging with thermal infrared (TIR) cameras can also be applied to
measure surface temperatures across large spatial scales. TIR cameras can also be
installed on unmanned aerial vehicles allowing for a landscape-level mapping at a
centimeter resolution. Accurate surface temperature measurements with TIR
imagery are to date still challenging because surface temperatures may still deviate
from actual microclimatic air temperatures experienced by organisms or because not
all surfaces have a similar thermal emissivity.

Finally, another approach to generate microclimate maps is to downscale


macroclimatic grids such as WorldClim (Fick and Hijmans 2017) with
high-resolution remote sensing data layers such as digital terrain models or canopy
height models. An important advantage of this technique is that these models are
based on long-term series of macroclimatic data, thus allowing predictions of how
microclimatic conditions vary across time (Zellweger et al. 2019). This is
particularly relevant for forests where understory plant community changes are
primarily driven by long-term microclimatic dynamics rather than the rate of
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
268 Historical Ecology

macroclimatic warming (Zellweger et al. 2020). Mechanistic microclimate models


offer an alternative method to downscale macroclimatic data based on a set of
abiotic variables such as longwave and shortwave radiation, air temperature, rainfall,
wind speed, relative humidity, elevation, coastal effects and canopy shading
(Maclean et al. 2019).

20.3. Why do historical microclimates matter? Impacts on biodiversity


from the plot to landscape scale

When plot data are collected from surveys that record concurrent forest
microclimatic conditions and understory species composition (not only of plants, but
also, for example, arthropods, fungi and soil bacteria) at the exact same location and
period, impacts of local forest microclimate change can be linked to population,
community and ecosystem change. For instance, species-specific response curves
along a given microclimatic parameter (e.g. thermal response curves) can be fitted to
compute indices of community weighted means, such as floristic temperatures, and,
when expressed over time, calculate the increasing dominance of warm-adapted
species, i.e. thermophilization. This provides interesting new avenues for ecological
assessments based on floristic temperatures. Finally, also herbarium specimens with
accurate collection coordinates offer novel opportunities to quantify impacts of
global change on understory plants (e.g. on phenology), if they can be linked to
historical microclimate data.

A study by Gorissen et al. (2004), which investigated the decline in woodland


butterflies in Flanders (Belgium), illustrates the value of creating more thermal
hotspots in forest stands. They used historical butterfly data from two woodlands in
Belgium to relate the changes in butterfly diversity to changes in management
practices. They found that almost all typical woodland butterflies and also species
that make use of open habitats within woodlands went extinct in the second half of
the 20th century. They linked these declines to afforestation of open areas and
management change towards high forest practices. In the past, these forests were
managed as coppice or CWS which allowed more light to enter the forests. They
therefore argued that creating more open conditions such as larger ridges, more open
area, may benefit woodland butterflies, and also thermophilous taxa from other
groups. However, creating more open habitat may result in reduced temperature
buffering capacity of tree canopies. The cool and moist microclimate that closed
canopies induce is essential for many forest organisms adapted to lower
temperatures, moist and shaded conditions, such as slugs, ectomycorrhizal and
wood-decaying fungi, mosses and other cool-adapted plants, moths and specific
groups of arthropods (fungivorous beetles, woodlice, etc.) (Figure 20.2). Many
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Forest Microclimates 269

forests in Europe have become denser and cooler due to an overall reduction in
forest management intensity during the past decades, which is beneficial for these
forest specialist organisms. However, if forest cover will be reduced or canopies
become more open in the future due to changes in disturbance or management, these
organisms are likely to be negatively affected by increased edge effects.

Studying microclimates at the landscape scale can significantly contribute to


understanding biodiversity and species distribution patterns. Landscape factors, such
as forest cover and urbanization, can mediate the responses of wintering birds to
winter temperatures through microclimate variability, access to resources or both
(Frey et al. 2016b). Microclimate temperatures can have a large effect on the
occupancy dynamics of forest birds in a montane forest landscape with high
microclimatic variability. This suggests that the microclimate-associated distribution
patterns reflect the species’ potential for behavioral buffering from climate change.
Also, for plants, diversity and distribution patterns can be linked to microclimatic
differences and connectivity at the landscape scale. Local habitat features of
hedgerows, including the microclimate, can shape plant diversity and composition
patterns, affecting the potential of hedgerows to function as refugium or corridor for
forest plant specialists (Vanneste et al. 2020). The increasing possibilities to map
and model forest microclimates and its drivers at the landscape scale will provide
species distribution models and biodiversity studies with important information to
gain more insight in historical species’ distributions and future developments under
different global change scenarios.

Figure 20.2. The butterfly Apatura iris (left) and the woodlice Ligidium hypnorum
(right) as representatives of, respectively, hot- and cold-spots in forests. Photos:
Wiske Teugels and Spinicornis-Gert Arijs. For a color version of this figure, see
www.iste.co.uk/decocq/ecology.zip
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
270 Historical Ecology

20.4. Conclusion

Forest microclimates have changed during the past century at the plot, forest and
landscape level. These changes are still ongoing due to drivers such as climate
warming, changes in forest management and land-use. In many forests, the degree of
microclimatic buffering has increased due to a change from coppice-based systems
to high forest. But at the same time, buffering along (external) edges has decreased
over time due to increasing fragmentation, affecting the forest interior microclimate.
To cater to the needs of forest organisms with contrasting microclimatic
requirements in a changing climate, several management options are available. For
instance, specific zoning systems can be developed to create both hot and cold spots
in forested landscapes, considering the surrounding landscape matrix. In this
endeavor, the historical knowledge of microclimates provides an invaluable source
of information to build on.

20.5. References

Arroyo-Rodríguez, V., Saldana-Vazquez, R.A., Fahrig, L., Santos, B.A. (2017). Does forest
fragmentation cause an increase in forest temperature? Ecol. Res., 32(1), 81–88.
Chytrý, M., Hennekens, S.M., Jiménez-Alfaro, B., Knollová, I., Dengler, J., Jansen, F.,
Landucci, F., Schaminée, J.H.J., Aćić, S., Agrillo, E. et al. (2016). European Vegetation
Archive (EVA): An integrated database of European vegetation plots. Appl. Veg. Sci.,
19(1), 173–180.
De Frenne, P., Zellweger, F., Rodríguez-Sánchez, F., Scheffers, B.R., Hylander, K., Luoto,
M., Vellend, M., Verheyen, K., Lenoir, J. (2019). Global buffering of temperatures under
forest canopies. Nat. Ecol. Evol., 3, 744–749.
De Frenne, P., Lenoir, J., Luoto, M., Scheffers, B.R., Zellweger, F., Aalto, J., Ashcroft, M.B.,
Christiansen, D.M., Decocq, G., De Pauw, K. et al. (2021). Forest microclimates and
climate change: Importance, drivers and future research agenda. Glob. Chang. Biol.,
27(11), 2279–2297
Fick, S.E. and Hijmans, R.J. (2017). Worldclim 2: New 1-km spatial resolution climate
surfaces for global land areas. Int. J. Climatol., 37(12), 4302–4315.
Frey, S.J.K., Hadley, A.S., Johnson, S.L., Schulze, M., Jones, J.A., Betts, M.G. (2016a).
Spatial models reveal the microclimatic buffering capacity of old-growth forests. Sci.
Adv., 2, e1501392.
Frey, S.J.K., Hadley, A.S., Betts, M.G. (2016b). Microclimate predicts within-season
distribution dynamics of montane forest birds. Divers. Distrib., 22, 944–959.
Geiger, R., Aron, R.H., Todhunter, P. (2003). The Climate Near the Ground. Rowman and
Littlefield, Oxford, UK.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Forest Microclimates 271

Gorissen, D., Merckx, T., Vercoutere, B., Maes, D. (2004). Veranderd bosgebruik en
dagvlinders. Landschap, 21, 85–95.
Maclean, I.M.D., Mosedale, J.R., Bennie, J.J. (2019). Microclima: An R package for
modelling meso- and microclimate. Methods Ecol. Evol., 10(2), 280–290.
den Ouden, J., Muys, B., Mohren, F., Verheyen, K. (2010). Bosecologie en Bosbeheer. Acco,
Leuven, Belgium.
Szabó, P. (2015). Historical ecology: Past, present and future. Biol. Rev., 90, 997–1014.
Tymen, B., Vincent, G., Courtois, E.A., Heurtebize, J., Dauzat, J., Marechaux, I., Chave, J.
(2017). Quantifying micro-environmental variation in tropical rainforest understory at
landscape scale by combining airborne LiDAR scanning and a sensor network. Ann. For.
Sci., 74, 32.
Vanneste, T., Govaert, S., Spicher, F., Brunet, J., Cousins, S.A.O., Decocq, G., Diekmann, M.,
Graae, B.J., Hedwall, P., Kapás, R.E. et al. (2020). Contrasting microclimates among
hedgerows and woodlands across temperate Europe. Agric. For. Meteorol., 281, 107818.
Verheyen, K., De Frenne, P., Baeten, L., Waller, D.M., Hédl, R., Perring, M.P., Blondeel, H.,
Brunet, J., Chudomelová, M., Decocq, G. et al. (2017). Combining biodiversity resurveys
across regions to advance global change research. BioScience, 67(1), 73–83.
Zellweger, F., De Frenne, P., Lenoir, J., Rocchini, D., Coomes. D. (2019). Advances in
microclimate ecology arising from remote sensing. Trends Ecol. Evol., 34, 327–341.
Zellweger, F., De Frenne, P., Lenoir, J., Vangansbeke, P., Verheyen, K. et al. (2020). Forest
microclimate dynamics drive plant responses to warming. Science, 368(6492), 772–775.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
21

Causes and Consequences of


Extinction Debts: Perspectives
for Historical Ecology and
Biological Conservation
Grégoire BLANCHARD1 and François MUNOZ2
1
University of Montpellier, France
2
Grenoble Alpes University, Grenoble, France

21.1. Introduction

Among the concepts that are pivotal in historical ecology, extinction debt has
received considerable attention and strong support from both theoretical and
empirical evidence (Malanson 2008; Figueiredo et al. 2019). A broad definition of
extinction debt is that, after a modification of ecosystem characteristics, we can
observe populations of some species at a given time and place although they are
doomed to become extinct under current ecological conditions (Kuussaari et al.
2009). Such extinction delay can last over decades or centuries (Hylander and
Ehrlén 2013). Assuming that today’s biodiversity patterns represent a snapshot of
ecosystems at equilibrium can then be misleading for conservation planning, making
uncertainty on the causes and consequences of these long transient dynamics a
crucial concern for biodiversity research (Kuussaari et al. 2009). While
anthropogenic alteration of ecosystems becomes faster and stronger, there is a
growing need to address the diversity of ecological – and historical – scenarios
leading to extinction debts, as well as implications and opportunities for biological
conservation (Figueiredo et al. 2019).

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
274 Historical Ecology

Accumulating long-term, empirical data on ecosystem history sheds light on the


mechanisms generating delayed extinctions (Helm et al. 2005; Vellend et al. 2006;
Vellend et al. 2013). An illustrative example of the usefulness of historical ecology
in understanding the causes and processes generating extinction debt is the history of
North American National Parks. At the end of the 19th century, several large nature
reserves were settled in western North America. Despite strict conservation policies,
a critical decline of mammal populations was observed within parks during the 20th
century (Newmark 1995). Habitat fragmentation due to habitat destruction outside
the Parks entailed the progressive loss of biodiversity, consistent with the dynamics
predicted by the theory of island biogeography (MacArthur and Wilson 1967;
Newmark 1995). Specifically, if protected areas are too small for long-term
persistence of local populations and without immigration from outside the parks, a
progressive decline of biodiversity is expected (Newmark 1995). Similarly, there is
growing evidence for long time lags between expected and observed plant
extinctions following anthropogenic landscape changes (Vellend et al. 2006; Cronk
2016). Historical ecology thus provides valuable opportunities for studying
extinction debts and underlying processes, through taking advantage of long-term
and large-scale empirical scenarios from the past.

In this chapter, we expose ecological mechanisms leading to extinction debts,


and address implications for historical ecology from both research and conservation
perspectives. We discuss how, by shedding light on past ecological dynamics,
historical ecology is pivotal to detect and understand the mechanisms causing
extinction debts. Hindcasting past extinction debts through the prism of ecosystem
history can provide new insights and opportunities for future biodiversity
conservation.

21.2. Causes and processes entailing extinction debts

Extinction debts arise from different mechanisms and under diverse historical
scenarios (Hylander and Ehrlén 2013). Here, we propose a brief synthesis of the
historical causes and ecological processes underlying extinction debts, showing how
changes in basic properties of habitat (quantity, connectivity and quality) and of
biotic interactions can trigger delayed extinctions (Figure 21.1).

The persistence of species populations depends on local environmental


conditions, which represent the “habitat quality” for each species. Drastic alteration
of habitat quality can lead to rapid population extinction, but smaller or more
gradual changes can entail slow population decline (Hylander and Ehrlén 2013).
Historical changes in habitat quality can generate delayed extinctions in areas
progressively becoming unsuitable for population persistence. For instance,
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Causes and Consequences of Extinction Debts 275

human-driven climate warming since the mid 20th century causes delayed plant
species extinction in European Alps (Rumpf et al. 2019). Changing disturbance
regimes (e.g. land-use practices, wildfires) can also affect habitat quality and result
in extinction debts (Essl et al. 2015). Importantly, long-lived species can be subject
to lags in demographic decline, notably if environmental change affects only
specific life stages (e.g. reproduction or recruitment, Hylander and Ehrlén (2013)).
For instance, individual trees may persist for hundreds of years even though low
habitat quality prevents seedling recruitment (Vellend et al. 2006). Conversely,
changes in environmental conditions can also generate an immigration credit for
species adapted to novel environmental conditions (Box 21.1).

Figure 21.1. Forcing events (historical causes) and temporal biodiversity trajectory
associated with an extinction debt. (a) Red arrows indicate changes in abiotic (habitat
characteristics) and biotic (species extinction or invasion) ecosystem conditions,
which entail extinction debts. (b) Extinction debt is progressively paid during the
relaxation time following the forcing events in (a), until a new equilibrium is reached.
New environmental conditions in (a) can allow establishment of new immigrating
species, but the immigration can also be delayed, resulting in an immigration credit
(Box 1). Biodiversity change (grey area) depends on the net balance (grey dotted
line) between extinctions (red line) and immigration (blue dotted line). For a color
version of this figure, see www.iste.co.uk/decocq/ecology.zip

Habitat loss and fragmentation are further drivers of extinction debt, which
received most interest (Tilman et al. 1994). The available quantity of a given habitat
directly determines the quantity of resources for species, and hence, the size of
populations (MacArthur and Wilson 1967). Following past habitat loss, smaller
remnant populations entail greater demographic stochasticity, increasing the risk of
extinctions (Hylander and Ehrlén 2013). Habitat connectivity determines the ability
of species to immigrate in habitat patches. Thus, lower connectivity resulting from
past habitat fragmentation reduces immigration rates, and increases the probability
of population extinction (Essl et al. 2015). Historical decreases in habitat quantity
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
276 Historical Ecology

and/or connectivity are likely to push multiple species toward stochastic extinction,
triggering large and long-lasting extinction debts (Hylander and Ehrlén 2013).

At an equilibrium of extinction and immigration dynamics, biodiversity is expected to


reach stationarity. Extinction debt is a transient biodiversity state during which species
loss is delayed after a forcing event altering the initial ecosystem conditions (Figure 21.1).
Conversely, the changes in abiotic or biotic ecosystem conditions can create opportunities
for other species to immigrate. Because it can take time for these species to colonize new
areas, immigration can be delayed as extinction can be, entailing an “immigration credit”
(Jackson and Sax 2010). The “net” biodiversity change (i.e. loss or gain of species)
depends on the difference between immigration credit and extinction debt (Figure 21.1).
Immigration credit can be paid faster or slower than extinction debt is, generating
transient biodiversity excess or deficit, respectively. The relaxation time represents the
amount of time during which extinction debt and/or immigration credit are paid until
biodiversity dynamics reaches the new equilibrium (Kuussaari et al. 2009).

Box 21.1. Extinction debt, immigration credit and relaxation time

Changes in biotic ecosystem context can also cause extinction debts (Mouquet
et al. 2011; Figueiredo et al. 2019). For instance, the invasion by generalist and
strong competitors, or by a new predator, can entail gradual extinction of native
species (Didham et al. 2007; Gilbert and Levine 2013). These are classical scenarios
on oceanic islands where many human-driven species invasions have generated
extinction debts lasting for decades or centuries (Sax and Gaines 2008; Cronk 2016).
Moreover, the extinction of a given species can entail the delayed co-extinction of
other species through loss and disruption of interactions, generating cascading
extinction debt (Valiente‐Banuet et al. 2015). Delayed co-extinctions are particularly
expected to arise following the extinction of keystone or engineer species having
key roles in ecosystem functioning (Estes et al. 2011). A striking example of such
delayed co-extinction dynamics is the slow decline of large-seeded tree species
resulting from long-term human-driven defaunation of large-bodied dispersers in
tropical forests (Terborgh et al. 2008).

21.3. Studying and detecting extinction debts from ecosystem history

Research challenges arising from extinction debts stand in the long-term delay
between their historical causes and ecological consequences (Kuussaari et al. 2009).
Extinction debts cannot be studied from snapshots of current ecosystem conditions,
and often lag over timescales too long to match the duration of most ecological
monitoring. Available data from the past, as provided by a variety of sources (see
Vellend et al. 2013), are thus essential to investigate the trajectory of ecosystem
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Causes and Consequences of Extinction Debts 277

dynamics and deduce the legacies of past ecosystem conditions and historical
changes (Kuussaari et al. 2009; Figueiredo et al. 2019). Historical ecology is thus at
the forefront of establishing causal links between past events and delayed extinction
dynamics (Figure 21.2).

Figure 21.2. Synthetic framework for studying, detecting and mitigating extinction
debts. Information about ecosystem history helps to detect extinction debt and
understand underlying mechanisms. Then, adequate conservation and management
measures can be developed to mitigate extinction debts. Conservation perspectives
focusing on ecosystem functioning should better succeed in maintaining or restoring
ecosystem biodiversity and services than taxon- and habitat-focused approaches.
For a color version of this figure, see www.iste.co.uk/decocq/ecology.zip

Landscape ecology probably encompasses most of the empirical approaches used


to assess delayed extinction dynamics (Box 21.2, Figure 21.3). Landscape history
can be assessed through terrestrial and aerial historical photographs, as well as from
historical maps (Vellend et al. 2013). Even when they are not related to a change in
land-cover, detailed information about past land-use practices (e.g. agricultural and
pastoral practices, soil fertilization, logging and forest management practices,
hunting, stream water use) can also provide important insights about past habitat
characteristics and ecosystem legacies (Vellend et al. 2013; Bürgi et al. 2017;
Bergès and Dupouey 2020). Although only available for the last few decades,
satellite images are also a promising source of information about landscape history
for investigating extinction debts at the regional scale (Jung et al. 2019). Analyzing
how species with different functional traits respond to recent changes in ecosystem
conditions should help to identify the specific factors of extinction debts (Hylander
and Ehrlén 2013). Such approaches can be used to infer parameters of mechanistic
community assembly models (Blanchard et al. 2020), providing a more explicit
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
278 Historical Ecology

understanding of the ecological mechanisms at play, and enabling simulations of


future ecosystem dynamics (Box 21.2).

Detecting extinction debts can be achieved in different ways, depending on the


available data describing past and present biodiversity and ecosystem conditions. Most
approaches assume that past ecosystem immigration–extinction dynamics were at
equilibrium before a given change in ecosystem conditions (Kuussaari et al. 2009).
Historical data can thus be used to estimate relationships linking biodiversity and
ecosystem abiotic and/or biotic conditions (e.g. habitat quality, quantity, connectivity,
abundance of keystone species or invasive predator) at equilibrium. Then, comparing
present-day biodiversity and ecosystem conditions with the expected relationship at
equilibrium makes it possible to detect potential extinction debts (Figure 21.3). When no
data about past biodiversity is available, a widespread method is to use species–area
relationships to evaluate if current biodiversity patterns match better with past than
current habitat quantity and connectivity, to reveal ongoing extinction debt (Helm et al.
2005). If no historical data are available at all for the landscape of interest, extinction debt
can also be detected by comparing present-day biodiversity patterns and ecosystem
conditions with those from another similar landscape having a more stable history
(Vellend et al. 2006). However, these methods do not provide direct estimations.
Explicitly quantifying extinction debts (i.e. the expected number of future extinctions) is
possible if past and present data are available for both biodiversity and ecosystem
conditions. The combination of past and present data can be used to parameterize
mechanistic models to assess current and future ecosystem dynamics according to varying
landscape scenarios (May et al. 2013; Blanchard et al. 2020). Such data can also be used
to fit species distribution models in order to evaluate and predict lags in species extinction
and colonization following environmental changes (Blois et al. 2013; Talluto et al. 2017).

Box 21.2. Detecting and predicting extinction debts from historical data

Linking past ecosystem conditions and current biodiversity can still fail to
unambiguously detect extinction debts (see Box 21.2), and, while rarely available,
only long-term monitoring of biodiversity makes it possible to explicitly track
species extinction dynamics (Kuussaari et al. 2009). For instance, Jones et al. (2018)
illustrate how decades of time-series data makes it possible to relate local species
extinction dynamics to landscape management history. Alternative sources of
historical data also help to reconstruct past trajectories of biodiversity patterns and
dynamics over longer periods (Vellend et al. 2013). Natural history collections offer
many opportunities to assess past composition of ecosystems (Johnson et al. 2011).
For instance, most herbarium collections contain records spanning over the entire
20th century and beyond, which can be used to evaluate past geographical
distributions and phenology of species, or to estimate the dates of species invasions
or extinction (Sax and Gaines 2008; Johnson et al. 2011; Vellend et al. 2013). Over
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Causes and Consequences of Extinction Debts 279

larger time-scales, palynological, paleoecological records can help to track past


changes in biodiversity and land cover in relation to human influences (Vellend
et al. 2013; Bürgi et al. 2017). Such data can serve both to estimate past biodiversity
state and past biotic conditions (e.g. abundance of keystone species) of ecosystems.

Figure 21.3. General approach for evaluating extinction debt from landscape history.
In the example here, a given landscape with six habitat patches is subject to recent
changes in patch-level conditions (e.g. habitat loss). The relationship between
biodiversity and ecosystem conditions at equilibrium is deduced from historical data
(see Box 21.2). This relationship is then used to estimate extinction debt following
changes in ecosystem conditions. The mismatch between this relationship at
equilibrium and current biodiversity and ecosystem conditions (e.g. patch area)
indicates an extinction debt. The amplitude of extinction debts associated with future
fragmentation scenarios can be forecasted accordingly. For a color version of this
figure, see www.iste.co.uk/decocq/ecology.zip

Inertia in species extinction dynamics along with biotic dependency across


species can entail co-extinction debts. Historical data provide valuable information
about how changes in ecosystem conditions trigger delayed co-extinction. Numerous
historical cases of co-extinction have been documented (Estes et al. 2011; Essl et al.
2015), including ancient trophic collapses (Roopnarine 2006). For instance,
paleolimnology provides helpful quantitative tools to evaluate how past
environmental changes have affected the structure of trophic networks in freshwater
ecosystems (Jeppesen et al. 2001). These data can help to understand which
combination of environmental change, characteristics of species (e.g. specific traits
and trophic position) and ecosystem properties (food web structure, functional
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
280 Historical Ecology

diversity), can entail co-extinction dynamics (Pillai et al. 2010; Brodie et al. 2014).
Conclusions drawn from historically documented ecosystems where extinction debt
has been partly or totally paid should be used, albeit with care, to make generalized
predictions for similar and more recently affected systems. Notably, past ecosystem
history can help to identify early warning signals of incipient critical transitions
associated with pervasive extinction debts (Essl et al. 2015). Integrating more and
more historical data from different sources should provide new opportunities to
improve our understanding of how co-extinction debts relate to ecosystem legacies
(Bürgi et al. 2017).

21.4. Implications for biodiversity conservation and management

While extinction debts are undoubtedly a challenge for biodiversity


conservation, their delayed nature can let enough time to take measures preventing,
or at least mitigating, future losses of biodiversity and ecosystem services (Figure
21.2, Vellend et al. 2006). This can be achieved through rapid and relevant
restoration of habitat characteristics and ecosystem functions, before the debt is paid
(Hylander and Ehrlén 2013; Cronk 2016). In addition, habitat areas with long-term
environmental stability have high conservation values, as they are likely to act as
refugia for remnant populations, and so as biodiversity reservoirs for future
ecosystem restoration (Box 21.3). Increasing habitat quantity and connectivity
should help to prevent the extinction of remnant populations. This can be done by
identifying most spatially and environmentally suitable areas for habitat restoration
(Newmark et al. 2017). However, the effects of restoration actions are likely to be
delayed via an immigration credit, and this should be considered in conservation
planning (Jackson and Sax 2010; Cronk 2016). When extinction debt arises from
decreasing habitat quality, solutions should focus on maintaining or improving
environmental conditions, and reducing stress factors. In the context of climate
change, densifying vegetation canopy and reducing grazing pressure can help to
maintain plant species populations (Vaughan and Gotelli 2021). Favoring species
regional persistence by preserving and restoring thermally buffered areas (e.g.
topographic concavities) might also mitigate climatic extinction debts (Lenoir et al.
2017).

Biological conservation is still mostly focused on preserving or improving


habitat quality and spatial structure to reduce species extinction (Brodie et al. 2018).
Detecting extinction debts and scheduling conservation action for all species
interacting in an ecosystem is difficult, if not impossible, and only a few most
endangered species are generally prioritized. Switching towards ecosystem-focused
conservation policies with a more explicit integration of species function and
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Causes and Consequences of Extinction Debts 281

interactions should help to mitigate delayed (co-)extinction dynamics (Figure 21.2).


Identifying and conserving species providing key and unique ecological functions
should have substantial positive effects in maintaining or restoring ecosystem-level
biodiversity and services (Brodie et al. 2014, 2018, see also Box 21.3). Here again,
historical ecology can serve to retrospectively evaluate how biodiversity legacies
relate to past changes in ecosystem functioning (Ordonez and Svenning 2020), and
which management practices have succeeded or failed in maintaining essential
ecosystem functions (Figure 21.2).

European and North American forests have been subject to important human-driven
impacts during the last centuries resulting from land-use changes (e.g. logging, expansion
or abandonment of agricultural areas). Because they rely on slow demography of trees and
associated soil dynamics, their long-term dynamics is intimately linked to historical
anthropogenic influence (Foster et al. 2003; Hermy and Verheyen 2007). The abundance
and quality of long-term historical data on these forests revealed that they are particularly
heterogeneous in terms of age and historical legacies. As a result, they generally undergo
long-term transient dynamics associated with both extinction debts and immigration
credits (Vellend et al. 2006; Bergès and Dupouey 2020). The study of old forests (i.e.
forests with long-term stable conditions) versus recent forests provides valuable insights
for understanding the links between ecosystem history and delayed ecological dynamics,
which should guide adequate conservation measures (Hermy and Verheyen 2007). Old
forest patches have a high conservation value because they contain an important pool of
forest specialist plants that slowly colonize recently reforested areas, entailing an
immigration credit in recent forests (Hermy and Verheyen 2007; Bergès and Dupouey
2020). Moreover, large old trees often act as ecosystem engineers that provide resources,
microclimatic conditions and physical habitat for many other plant and animal species
(Jones et al. 2018). Preserving forest areas containing high density of old trees should thus
sustain trophic networks and ecosystem functions.

Box 21.3. The case of old forests

21.5. Conclusion

The influence of extinction debts is pervasive under rapidly changing


environmental conditions and landscape structures. However, conservation baselines
are generally determined from short-term socio-ecological contexts and issues that
are mostly blind to longer-term and transient ecological dynamics (Crumley et al.
2017). It is thus crucial to build conservation and land management policies that
better integrate the history and legacies of ecosystems (Bürgi et al. 2017; Crumley
et al. 2017). Achieving this goal will require both the future development of
multidisciplinary research frameworks and raising awareness of the inertia driving
ecosystem dynamics, whereby historical ecology should play a key role.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
282 Historical Ecology

21.6. References

Bergès, L. and Dupouey, J. (2020). Historical ecology and ancient forests: Progress,
conservation issues and scientific prospects, with some examples from the French case.
J. Veg. Sci., 32(1).
Blanchard, G., Birnbaum, P., Munoz, F. (2020). Extinction–immigration dynamics lag behind
environmental filtering in shaping the composition of tropical dry forests within a
changing landscape. Ecography, 43(6), 869–881.
Blois, J.L., Zarnetske, P.L., Fitzpatrick, M.C., Finnegan, S. (2013). Climate change and the
past, present, and future of biotic interactions. Science, 341(6145), 499–504.
Brodie, J.F., Aslan, C.E., Rogers, H.S., Redford, K.H., Maron, J.L., Bronstein, J.L., Groves,
C.R. (2014). Secondary extinctions of biodiversity. Trends Ecol. Evol., 29(12), 664–672.
Brodie, J.F., Redford, K.H., Doak, D.F. (2018). Ecological function analysis: Incorporating
species roles into conservation. Trends Ecol. Evol., 33(11), 840–850.
Bürgi, M., Östlund, L., Mladenoff, D.J. (2017). Legacy effects of human land use:
Ecosystems as time-lagged systems, Ecosystems, 20(1), 94–103.
Cronk, Q. (2016). Plant extinctions take time. Science, 353(6298), 446–447.
Crumley, C.L., Lennartsson, T., Westin, A. (eds) (2017). Issues and Concepts in Historical
Ecology: The Past and Future of Landscapes and Regions, 1st edition. Cambridge
University Press, Cambridge.
Didham, R., Tylianakis, J., Gemmell, N., Rand, T., Ewers, R. (2007). Interactive effects of
habitat modification and species invasion on native species decline. Trends Ecol. Evol.,
22(9), 489–496.
Essl, F., Dullinger, S., Rabitsch, W., Hulme, P.E., Pyšek, P., Wilson, J.R.U., Richardson,
D.M. (2015). Delayed biodiversity change: No time to waste. Trends Ecol. Evol., 30(7),
375–378.
Estes, J.A., Terborgh, J., Brashares, J.S., Power, M.E., Berger, J., Bond, W.J., Carpenter,
S.R., Essington, T.E., Holt, R.D., Jackson, J.B.C. et al. (2011). Trophic downgrading of
planet earth. Science, 333(6040), 301–306.
Figueiredo, L., Krauss, J., Steffan-Dewenter, I., Sarmento Cabral, J. (2019). Understanding
extinction debts: Spatio–temporal scales, mechanisms and a roadmap for future research.
Ecography, ecog.04740.
Foster, D., Swanson, F., Aber, J., Burke, I., Brokaw, N., Tilman, D., Knapp, A. (2003). The
importance of land-use legacies to ecology and conservation. BioScience, 53(1), 77–88.
Gilbert, B. and Levine, J.M. (2013). Plant invasions and extinction debts. Proc. Natl. Acad.
Sci., 110(5), 1744–1749.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Causes and Consequences of Extinction Debts 283

Helm, A., Hanski, I., Partel, M. (2005). Slow response of plant species richness to habitat loss
and fragmentation. Ecol. Lett., 9(1), 72–77.
Hermy, M. and Verheyen, K. (2007). Legacies of the past in the present-day forest
biodiversity: A review of past land-use effects on forest plant species composition and
diversity. Ecol. Res., 22(3), 361–371.
Hylander, K. and Ehrlén, J. (2013). The mechanisms causing extinction debts. Trends Ecol.
Evol., 28(6), 341–346.
Jackson, S.T. and Sax, D.F. (2010). Balancing biodiversity in a changing environment:
Extinction debt, immigration credit and species turnover. Trends Ecol. Evol., 25(3),
153–160.
Jeppesen, E., Leavitt, P., De Meester, L., Jensen, J.P. (2001). Functional ecology and
palaeolimnology: Using cladoceran remains to reconstruct anthropogenic impact. Trends
Ecol. Evol., 16(4), 191–198.
Johnson, K.G., Brooks, S.J., Fenberg, P.B., Glover, A.G., James, K.E., Lister, A.M., Michel,
E., Spencer, M., Todd, J.A., Valsami-Jones, E. et al. (2011). Climate change and
biosphere response: Unlocking the collections vault. BioScience, 61(2), 147–153.
Jones, G.M., Keane, J.J., Gutiérrez, R.J., Peery, M.Z. (2018). Declining old-forest species as a
legacy of large trees lost. Divers. Distrib., 24(3), 341–351.
Jung, M., Rowhani, P., Scharlemann, J.P.W. (2019) Impacts of past abrupt land change on
local biodiversity globally, Nat. Commun., 10(1), 5474.
Kuussaari, M., Bommarco, R., Heikkinen, R.K., Helm, A., Krauss, J., Lindborg, R., Öckinger,
E., Pärtel, M., Pino, J., Rodà, F. et al. (2009). Extinction debt: A challenge for
biodiversity conservation. Trends Ecol. Evol., 24(10), 564–571.
Lenoir, J., Hattab, T., Pierre, G. (2017). Climatic microrefugia under anthropogenic climate
change: Implications for species redistribution. Ecography, 40(2), 253–266.
MacArthur, R. and Wilson, E. (1967). The Theory of Island Biogeography. Princeton
University Press, Princeton.
Malanson, G.P. (2008) Extinction debt: Origins, developments, and applications of a
biogeographical trope. Prog. Phys. Geogr., 32(3), 277–291.
May, F., Giladi, I., Ristow, M., Ziv, Y., Jeltsch, F. (2013a). Metacommunity, mainland-island
system or island communities? Assessing the regional dynamics of plant communities in a
fragmented landscape. Ecography, 36(7), 842–853.
Mouquet, N., Matthiessen, B., Miller, T., Gonzalez, A. (2011). Extinction debt in source-sink
metacommunities. PLOS ONE, 6(3), e17567.
Newmark, W.D. (1995) Extinction of mammal populations in western north american
national parks. Conserv. Biol., 9(3), 512–526.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
284 Historical Ecology

Newmark, W.D., Jenkins, C.N., Pimm, S.L., McNeally, P.B., Halley, J.M. (2017). Targeted
habitat restoration can reduce extinction rates in fragmented forests. Proc. Natl. Acad.
Sci., 114(36), 9635–9640.
Ordonez, A. and Svenning, J.-C. (2020). The potential role of species and functional
composition in generating historical constraints on ecosystem processes. Glob. Ecol.
Biogeog., 29(2), 207–219.
Pillai, P., Loreau, M., Gonzalez, A. (2010). A patch-dynamic framework for food web
metacommunities. Theor. Ecol., 3(4), 223–237.
Roopnarine, P.D. (2006). Extinction cascades and catastrophe in ancient food webs.
Paleobiology, 32(1), 1–19.
Rumpf, S.B., Hülber, K., Wessely, J., Willner, W., Moser, D., Gattringer, A., Klonner, G.,
Zimmermann, N.E., Dullinger, S. (2019). Extinction debts and colonization credits of
non-forest plants in the European Alps. Nat. Commun., 10, 4293.
Sax, D.F. and Gaines, S.D. (2008). Species invasions and extinction: The future of native
biodiversity on islands. Proc. Natl. Acad. Sci., 105(Suppl. 1), 11490–11497.
Talluto, M.V., Boulangeat, I., Vissault, S., Thuiller, W., Gravel, D. (2017). Extinction debt
and colonization credit delay range shifts of eastern North American trees. Nat. Ecol.
Evol., 1(7), 0182.
Terborgh, J., Nuñez-Iturri, G., Pitman, N.C.A., Valverde, F.H.C., Alvarez, P., Swamy, V.,
Pringle, E.G., Paine, C.E.T. (2008). Tree recruitment in an empty forest. Ecology, 89(6),
1757–1768.
Tilman, D., May, R.M., Lehman, C.L., Nowak, M.A. (1994). Habitat destruction and the
extinction debt. Nature, 371(6492), 65–66.
Valiente-Banuet, A., Aizen, M.A., Alcántara, J.M., Arroyo, J., Cocucci, A., Galetti, M.,
García, M.B., García, D., Gómez, J.M., Jordano, P. et al. (2015). Beyond species loss: The
extinction of ecological interactions in a changing world. Funct. Ecol., 29(3), 99–307.
Vaughan, I.P. and Gotelli, N.J. (2021). Using climatic credits to pay the climatic debt. Trends
Ecol. Evol., 36(2), 104–112.
Vellend, M., Verheyen, K., Jacquemyn, H., Kolb, A., Calster, H.V., Peterken, G., Hermy, M.
(2006). Extinction debt of forest plants persists for more than a century following habitat
fragmentation. Ecology, 87(3), 542–548.
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. Am. J. Bot., 100(7), 1294–1305.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
22

Historical Ecology for the Past


and the Future: Organizing at
Local and Regional Scales
Carole L. CRUMLEY1,2
1
University of North Carolina, Chapel Hill, USA
2
Swedish University of Agricultural Sciences, Uppsala, Sweden

22.1. Introduction

Multiple unfolding crises menace not just humanity but all life on Earth.
National, international and global institutions are moving too slowly to make a
difference in the outcome for future generations. In the past two decades, the
inability to come to agreements – and keep them – has stalled multinational progress
towards goals that affect the entire planet (Clemençon 2016, Maizland 2021). This
impasse has expanded scholarly and public interest in initiatives at local and
regional scales; these can bring new energy and ideas to sluggish transnational
scales and influence broader constituencies (e.g. Raja and Simpson 2021).

The action list is long: increase the pool of expertise by engaging all relevant
knowledge communities, collect rapidly disappearing data, analyze with both familiar
and new methods and apply actionable science to policy and practice (Palmer 2012;
Beier et al. 2017). This enormously complex and urgent activity requires integrated
research frameworks with the flexibility to accommodate the global diversity of
places, peoples and processes, and the means to examine future options. Rather than
waiting for funds and instructions, easily adaptable on-the-ground research designs
with broad applicability can better guide practical efforts.

Historical Ecology,
coordinated by Guillaume DECOCQ. © ISTE Ltd 2022.
Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
286 Historical Ecology

The Integrated History and Future of People on Earth (IHOPE) is an early


example of how such initiatives by researchers have powered collaborations with
diverse stakeholders, including residents and citizen scientists, local and regional
administrative bodies, and commercial entities. The IHOPE example is particularly
useful because it has benefitted from members’ individual and stakeholder
collaboration. In this fashion, the motor of development of durable practice is in the
hands of those most knowledgeable of conditions, limitations and opportunities for
action.

22.2. Founding IHOPE

Umbrella international scientific organizations were not particularly common in


the 1980s. During its three decades of existence, the International Geosphere–
Biosphere Programme (IGBP, 1986–2015), to better understand the Earth system,
built research networks, facilitated synthesis and enhanced capacity around the
world. The IGBP led a global effort to study how planet Earth functions and how it
has – and will – change. The IGBP’s framework linked land, sea and sky projects
where they meet: land–atmosphere, land–ocean and atmosphere–ocean. The IGBP’s
clear global goals attracted 78 national committees, solidly representing every
continent, and supported in-country collaborations through conferences and
workshops; these encouraged both established researchers and new scientists to
frame their research to address global challenges (IGBP 2015).

The Global Change Open Science Conference held in Amsterdam in 2001 was
jointly organized by the IGBP, the International Human Dimensions Programme on
Global Environmental Change (IHDP), the Global Programme on Biodiversity
(DIVERSITAS) and the World Climate Research Programme (WCRP). The
conference produced the “Amsterdam Declaration”, which affirmed that the extent
and impact of anthropogenic forces are equal to those of nature and called for an
“ethical framework for global stewardship and strategies for Earth system
management” (Moore III et al. 2001). This marked the global first steps towards
including human activity in the analysis of Earth system research, a goal shared
across the spectrum of IGBP and IHDP global programs.

At an IGBP conference in Banff, Canada in 2003, ecological economist Robert


Costanza, a member of the IGBP task force Analysis, Integration and Modelling of
the Earth System (AIMES), conceived the idea of constructing a timeline for human
interaction with the Earth system. Will Steffen – an Earth system scientist and then
executive director of the IGBP – agreed, viewing the initiative as consonant with the
IGBP’s aim to provide a holistic understanding of planetary dynamics.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Ecology for the Past and the Future 287

Trained as an archaeologist and environmental anthropologist, and a member of


the IGBP community (PAGES, AIMES), I quickly learned about the idea, soon
named the Integrated History and Future of People on Earth (IHOPE). My
archaeology colleagues and I were thrilled at the prospect of collaborating on such a
timeline, as it aligned with our efforts to build a framework that would enable better
integration of the Earth system sciences with the social sciences and humanities.

Keenly aware that we could not simply slap together different datasets and that
modeling the past has its own challenges, our group began to expand in numbers and
expertise. A Dahlem Conference was held in 2005, resulting in an edited book with
contributions from a broad range of scholars and titled Sustainability or Collapse?
An Integrated History and Future of People on Earth (Costanza et al. 2007a). This
book, along with articles, meetings and workshops (including workshops at
NCEAS), described the modeling and scenario building exercise and the framework
for data integration (Costanza et al. 2007b, 2012; Heckbert et al. 2016). The IHOPE
project office was first established in the United States (2007–2009; Hibbard et al.
2010), before relocating to Sweden (2009–present). After the dissolution of the
IGBP in 2015, IHOPE became a global research project of a new global program,
Future Earth (https://futureearth.org/).

22.3. Integrating the social sciences and humanities

The discussion between those who model and those who dig – and some who do
both – was fascinating and at times contentious. The modelers wanted to begin their
work right away, but the archaeologists and historians wanted to first establish a
protocol for data collection and analysis that would smoothly meld the diverse
perspectives and sources of evidence. In the end, there was a collective effort to
identify drivers and monitoring tools that could render the archaeological record
legible.

Trained to work with colleagues from many different disciplines and with many
types of information, the archaeologists led the integrating exercise. As the Annales
historian Marc Bloch (1993 [first published in 1949]) puts it:

Very few scholars can boast that they are equally well equipped to
read critically a medieval charter, to correctly interpret the etymology
of placenames, to date unerringly the remains of prehistoric
habitations, and to analyze the plants characteristic of a pasture, a
field, or a moor. Without all these, however, how could one pretend to
write the history of land-use? (author’s translation).
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
288 Historical Ecology

Current archaeological practice is not only designed to assist collaboration


among differently trained researchers but, in some national traditions, to also engage
stakeholders in the area under study. This is particularly the case in North America
where, despite brutal colonial histories, many indigenous peoples remain and are
intimately tied to the land. North American archaeologists trained in anthropology
departments have, over the last century, slowly moved archaeological practice
towards ethical positions as regards indigenous rights (NAGPRA 1990) and greater
awareness of the contemporary context of the past.

The research framework for IHOPE thus provides a kind of intellectual “contact
zone” (Pratt 1991), in which diverse communities can exchange information about
the past and the present of a physical region and discuss various courses of action
for its future. The disciplines that have slowly moved together into this zone are
anthropology, archaeology, history, geography and ecology. As in any contact zone,
such exchanges require a strong commitment to dialogue, extensive translation and
significant efforts at coordination, but they are feasible (Jones et al. 2018;
Southerland-Wilson et al. 2019; Hicks 2021). In addition to a complex systems
approach and attention to scales of time and space, IHOPE includes hybrid fields of
study, such as bioarchaeology and critical cultural studies.

22.4. Historical ecology

Historical ecology is a cluster of concepts that offers a holistic, practical


perspective to the study of environmental change. It may be applied to spatial and
temporal frames at any resolution, finding particularly rich data sources at what is
loosely termed the “landscape” scale, where human activity and biophysical systems
interact and archaeological historical, and ethnographic records are plentiful. The
term assumes a definition of ecology that includes humans as a component of all
ecosystems, and a definition of history that encompasses the history of the Earth
system as well as the social and physical past of our species (Balée 2006; Crumley
2007, 2013, 2015, 2016; Crumley et al. 2018). Historical ecology is not simply the
study of the past, but the application of knowledge gained from its study to the
human future.

The term seems to have almost been “in the air” beginning later in the 20th
century (Deevey et al. 19791; Bilsky 1980; Brooks 1985). Working in France in the
mid-1970s, my colleagues and I struggled to give a name to the history of dynamic
environments, finally adopting historical ecology to characterize our study of
long-term regional change (Crumley and Marquardt 1987). Our subsequent

1. Deevey was using the term historical ecology as early as 1964.


Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Ecology for the Past and the Future 289

collaboration with ecologists, archaeologists and anthropologists at a 1990 School


for Advanced Research (SAR) workshop resulted in the publication of Historical
Ecology: Cultural Knowledge and Changing Landscapes (Crumley 1994).

Independent of our work, the anthropologist and ethnobotanist William Balée


and colleagues, working in South America, enriched traditional cultural ecology to
include a strong historical approach to botany. Balée gathered anthropologists,
historians and geographers in a 1994 workshop; the papers were published as
Advances in Historical Ecology (Balée 1998; see also Balée 2006; Balée and
Erickson 2006). Working together since the mid-1990s, Balée and Crumley have
furthered the approach with the creation of publishing venues for historical ecology,
at first with Columbia University Press and now with Routledge.

Historical ecology is predicated on the assumption that it is possible to construct


evidence-validated narratives of landscape transformation that has resulted from the
continual interaction of diverse human activities with changing environmental
conditions. Historical ecology perforce draws on a broad spectrum of concepts,
methods, theories and evidence, taken from the biological and physical sciences,
ecology, the social sciences and the humanities. Derived from the standard
procedures of various disciplines, independent datasets provide critical
“cross-checks” upon one another, revealing patterns of association with researchers
and often yielding further intriguing questions.

Due to the Earth’s enormous environmental and cultural diversity, no single


research strategy is possible. Instead, the flexibility of the historical ecology
“toolbox” encourages the matching of available kinds of information with local and
regional conditions (Meyer and Crumley 2011). For example, both tropical and
boreal research settings are rich in evidence of past conditions, but the tools for their
treatment are necessarily different, in response to pragmatic and strategic issues such
as preservation, funding, stakeholders and access.

Many of the participants in the 1990 SAR workshop became members of the
IHOPE Scientific Steering Committee, so it is not surprising that the IHOPE
collective embraced the historical ecology framework. Attracted by the possibilities
offered by this framework, individuals with textual, artistic, linguistic and other
expertise from the environmental humanities joined the group.

In the last decade, IHOPE has increased attention to several areas. Climate
history (Ljungqvist et al. 2021), modeling and scenario building (Scarborough and
Isendahl 2020; Scarborough 2021), landscape research (Crumley et al. 2017), and
the powerful approach of complex systems thinking (Sinclair et al. 2018) are only a
few of IHOPE contributions that advance understanding of landscapes, regions and
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
290 Historical Ecology

cultural change. Most recently, IHOPE has completed a book that examines ancient
urban regions and applies our findings to the future (Murphy and Crumley 2022).

22.5. Conclusion

The role of global programs has changed over the last 20 years, from facilitating
research in the Earth system sciences to facilitating their integration with long-term
human cognition, activity and behavior. In 2018, the International Council for
Science (ICSU) and the International Social Science Council (ISSC) merged, now
the International Science Council (ISC). Among the newest of international
sustainability programs is the UNESCO BRIDGES global coalition that aims to
integrate humanities, social science and other knowledges into research, education
and action for global sustainability “at local and territorial scales” (UNESCO 2019).
This on-the-ground inclusion of human activity as a driver of global change
underscores the importance of seeing the Earth as a complex system.

Strengthened with a more comprehensive assessment of the issues, we can better


understand and respond to crisis in multiple domains. However, contrasting global
goals and national endeavors and diminished national and international funding will
curtail the geographical reach of global programs. The unraveling of the Paris
Accord and other setbacks have accelerated the tendency to shift scales/arenas of
action from broad global goals to regional and local endeavors, such as those of
IHOPE.

The opportunity to meet and work at the global scale with researchers from
disciplines beyond one’s own was exciting and led to IHOPE’s founding. IHOPE’s
contribution is to facilitate work at the local and regional levels that can build
towards global goals. IHOPE is well placed to offer a smooth path among scales,
from historical global climates and environments to those to come.

There is no single solution to today’s challenges, because global natural, cultural


and historical diversity is too great. Instead, we must work from the bottom up: from
local and regional projects to broader scales, with flexible frameworks and a variety
of research tools. A good example is the necessary rethinking of one-size-fits-all
agricultural practices that use the same management tools across different
landscapes, biomes, soils, climates/microclimates and heritage. The historical
ecology toolbox, a complex systems approach, and the shifting needs and conditions
of the future call for on-the-ground flexibility.

This requires a collaboration between scholars and diverse stakeholders. How to


expand the research you do to the next village, county or countryside? Now is the
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Ecology for the Past and the Future 291

time to join with like-minded colleagues in your region. Find them at conferences,
contact those whose work inspires you, plan a project within your administrative
precinct(s), explore alliances with interested regional partners, find funding together
by increasing your networks. In addition to regional collaborations, the new
concepts and ideas that you have can be shared and expanded; this is how IHOPE
came to be.

You may say that would be nice but there is no funding. But for many of the last
18 years, IHOPE has operated pro bono, with no employees or website manager,
powered instead by a group of colleagues. With sporadic funding since the
dissolution of the IGBP, we have maintained the essentials, keeping a website
running since 2013, becoming affiliated with similar endeavors and using
colleagues’ extensive networks to build a vibrant community. IHOPE researchers,
using our framework, have been quite successful in finding funding for their own
research. We feature their work and that of others who offer their projects on the
website portal; the website itself is widely used as a teaching tool. Thanks to the
Internet, it is possible to build and maintain a coalition without extensive funding.

For IHOPE veterans, the shared activity of building a framework began with the
dialogue between archaeologists and natural scientists; for the archaeologists, the
historical ecology workshop at SAR gave us confidence that the IHOPE framework
could work in very different places. Your ideas and your energy can carry us into a
liveable future.

22.6. References

Balée, W. (ed.) (1998). Advances in Historical Ecology. Columbia University Press, New
York.
Balée, W. (2006). The research program of historical ecology. Annu. Rev. Anthropol., 35(5),
15–24.
Balée, W.L. and Erickson, C.L. (2006). Time and Complexity in Historical Ecology: Studies
in the Neotropical Lowlands. Columbia University Press, New York.
Beier, P., Hansen, L.J., Helbrecht, L., Behar, D. (2017). A how-to guide for coproduction of
actionable science. Conserv. Lett., 10(3), 288–296.
Bilsky, L.J. (ed.) (1980). Historical Ecology: Essays on Environment and Social Change.
Kennikat Press, Port Washington, NY.
Bloch, M. (1953). The Historian’s Craft. Vintage, New York.
Bloch, M. (1993). Apologie pour l’histoire ou métier d’historien. Armand Colin, Paris.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
292 Historical Ecology

Brooks, D.R. (1985). Historical ecology: A new approach to studying the evolution of
ecological associations. Ann. Missouri Bot. Garden, 72(4), 660–680.
Clémençon, R. (2016). The two sides of the Paris climate agreement: Dismal failure or
historic breakthrough? J. Env. Dev., 25(1), 3–24.
Costanza, R.L., Graumlich, L.J., Steffen, W. (eds) (2007a). Sustainability or Collapse? An
Integrated History and Future of People on Earth. Dahlem Workshop Reports # 96. MIT
Press, Cambridge, MA.
Costanza, R.L., Graumlich, L., Steffen, W., Crumley, C.L., Dearing, J., Hibbard, K.,
Leemans, R., Redman, C., Schimel, D. (2007b). Sustainability or collapse: What can we
learn from integrating the history of humans and the rest of nature? Ambio, 36, 522–527.
Costanza, R.L., van der Leeuw, S., Hibbard, K., Aulenbach, S., Brewer, S., Burek, M.,
Cornell, S., Crumley, C.L., Dearing, J., Folke, C. et al. (2012). Developing an integrated
history and future of people on Earth (IHOPE). Curr. Opin. Env. Sustain., 4, 106–114.
Crumley, C.L. (ed.) (1994). Historical Ecology: Cultural Knowledge and Changing
Landscapes. School of American Research, Santa Fe.
Crumley, C.L. (2007). Historical ecology: Integrated thinking at multiple temporal and spatial
scales. In The World System and the Earth System: Global Socio-environmental Change
and Sustainability Since the Neolithic, Hornborg, A. and Crumley, C.L. (eds). Left Coast
Press, Walnut Creek, CA.
Crumley, C.L. (2013). Historical ecology in archaeology. In Encyclopedia of Global
Archaeology, Smith, C. (ed.). Springer, New York
Crumley, C.L. (2015). New paths into the Anthropocene. In Handbook of Historical Ecology
and Applied Archaeology, Isendahl, C. and Stump, D. (eds). Oxford University Press,
Oxford.
Crumley, C.L. (2016). Historical ecology. In The Wiley Blackwell International Encyclopedia
of Anthropology, Callan, H. (ed.). Wiley Blackwell, New York.
Crumley, C.L. and Marquardt, W.H. (eds) (1987). Regional Dynamics: Burgundian
Landscapes in Historical Perspective. Academic Press, San Diego.
Crumley, C.L., Kolen, J.C.A., de Kleijn, M., van Manen, N. (2017). Studying long-term
changes in cultural landscapes: Outlines of a research framework and protocol. Landsc.
Res., 42(8), 880–890.
Crumley, C.L., Lennartsson, T., Westin, A. (eds) (2018). Issues and Concepts in Historical
Ecology: The Past and Future of Landscapes and Regions. Cambridge University Press,
Cambridge.
Deevey, E.S., Rice, D.S., Rice, P.M., Vaughan, H.H., Brenner, M., Flannery, M.S. (1979).
Mayan urbanism: Impact on a tropical karst environment. Science, 206, 298–306.
Heckbert, S., Costanza R.L., Parrott L. (2016). Achieving sustainable societies: Lessons from
modelling the ancient Maya. Solutions, 5(5), 55–64.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Historical Ecology for the Past and the Future 293

Hibbard, K.A., Costanza R., Crumley C., van der Leeuw S., Aulenbach, S., Dearing J.,
Morais J., Steffen, W., Yasuda, Y. (2010). Developing an integrated history and future of
people on earth (IHOPE): Research plan. Global Change. IGBP Report No. 59. IGBP
Secretariat, Stockholm.
Hicks, M. (2021). How archeology harms and helps indigenous resistance: Lessons from
Wet’suwet’en [Online]. Available at: https: //www.gc.cuny.edu/ [Accessed 20 March
2021].
IGBP (2015). National Committee List [Online]. Available at: http://www.igbp.net/people/
nationalcommittees/nationalcommitteelist.4.950c2fa1495db7081e183f5.html [Accessed
26 March 2021].
Jones, E.A., Westin, A., Madry, S., Murray, S., Moen, J., Tickner, A. (2018). How to
operationalise collaborative research. In Issues and Concepts in Historical Ecology: The
Past and Future of Landscapes and Regions, Crumley, C.L., Lennartsson, T., Westin, A.
(eds). Cambridge University Press, Cambridge.
Ljungqvist, F.C., Seim, A., Huhtamaa, H. (2021). Climate and society in European history.
Wires Clim. Change, 12, e691.
Maizland, L. (2021). Global climate agreements: Successes and failures [Online]. Available
at: https://www.cfr.org/backgrounder/paris-global-climate-change-agreements [Accessed
27 March 2021].
Meyer, W.J. and Crumley, C.L. (2011). Historical ecology: Using what works to cross the
divide. In Atlantic Europe in the First Millennium BC: Crossing the Divide, Moore, T.
and Armada, L. (eds). Oxford University Press, Oxford.
Moore III, B., Underdal, A., Lemke, P., Loreau, M. (2001). Amsterdam declaration on Earth
system science [Online]. Available at: http://www.igbp.net/about/history/2001amsterdamde
clarationonearthsystemscience.4.1b8ae20512db692f2a680001312.html [Accessed 7 April
2021].
Murphy, J.T. and Crumley, C.L. (eds) (2022). If the Past Teaches, What Does the Future
Learn? Ancient Urban Regions and the Durable Future. Research in Urbanism Series,
Delft School of Architecture, TU Delft OPEN Publishing, Delft.
NAGPRA (1990). The Native American Graves Protection and Repatriation Act (NAGPRA).
Pub. L. 101-601, 25 U.S.C. 3001 et seq., 104 Stat. 3048, is a United States Federal Law
enacted on November 16, 1990.
Palmer, M.A. (2012). Socioenvironmental sustainability and actionable science. BioScience,
62(1), 5–6.
Pratt, M. (1991). Arts of the contact zone. Profession, 1991, 33–40.
Raja, R. and Simpson, I.A. (2021). Scaling up and zooming in: Global history and
high-definition archaeology perspectives on the longue durée of urban–environmental
relations in Gerasa (Jerash, Jordan). J. Global Hist., 16(1), 1–20.
Scarborough, V.L. (2021). What ancient landscapes contribute to climate change. Econ.
Anthropol., 8, 161–167.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
294 Historical Ecology

Scarborough, V.L. and Isendahl, C. (2020). Distributed urban network systems in the tropical
archaeological record: Towards a model for urban sustainability in the era of climate
change. Anthropocene Rev., 7(3), 1–23.
Sinclair, P., Moen, J., Crumley, C.L. (2018). Historical ecology and the longue durée. In
Issues and Concepts in Historical Ecology: The Past and Future of Landscapes and
Regions, Crumley, C.L., Lennartsson, T., Westin, A. (eds). Cambridge University Press,
Cambridge.
Southerland-Wilson, D., Spice, A., Armstrong, C.G. (2019). Compliance archaeology fails
indigenous people in British Columbia: An example from Unist’ot’en territory. The
Midden, 49(1), 26–29.
UNESCO (2019). Toward the establishment of Bridges: Action to promote sustainability
science [Online]. Available at: https://en.unesco.org/news/toward-establishment-bridges-
action-promote-sustainability-science [Accessed 8 April 2021].
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
List of Authors

Vincent BALLAND Jérôme BURIDANT


Bourgogne-Franche-Comté University Jules Verne University of Picardie
Dijon Amiens
France France

Christophe BALTZINGER Clémence CHAUDRON


INRAE University of Lorraine
UR EFNO Nancy
Nogent-sur-Vernisson France
France
Roger COLY
Antoine BECKER-SCARPITTA University of Tours
University of Helsinki France
Finland
and Marco CONEDERA
Institute of Botany of the Czech Academy Swiss Federal Research Institute WSL
of Science Cadenazzo
Průhonice Switzerland
Czech Republic
Carole L. CRUMLEY
Grégoire BLANCHARD University of North Carolina
University of Montpellier Chapel Hill
France USA
and
Boris BRASSEUR Swedish University of Agricultural
Jules Verne University of Picardie Sciences
Amiens Uppsala
France Sweden

Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
296 Historical Ecology

Pieter DE FRENNE Elisabet FERNÁNDEZ-CABRERA


Ghent University Universidad de Las Palmas de
Belgium Gran Canaria
Canary Islands
Emiel DE LOMBAERDE Spain
Ghent University
Belgium Fabrizio FERRETTI
CREA Research Centre for Forestry
Karen DE PAUW and Wood
Ghent University Arezzo
Belgium Italy

Pallieter DE SMEDT Catherine FRUCHART


Ghent University Bourgogne-Franche-Comté University
Belgium Besançon
France
Guillaume DECOCQ
Jules Verne University of Picardie Emilie GALLET-MORON
Amiens Jules Verne University of Picardie
France Amiens
France
Francesca DI PIETRO
University of Tours Leví GARCÍA-ROMERO
France Universidad de Las Palmas de
Gran Canaria
Alexa DUFRAISSE Canary Islands
Muséum National d’Histoire Naturelle Spain
Paris
France Frank S. GILLIAM
University of West Florida
Ove ERIKSSON Pensacola
Stockholm University USA
Sweden
Olivier GIRARDCLOS
Damien ERTLEN Bourgogne-Franche-Comté University
University of Strasbourg Besançon
France France
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
List of Authors 297

Karen GREIG Patrik KREBS


University of Otago Swiss Federal Research Institute WSL
Dunedin Cadenazzo
New Zealand Switzerland

Luis HERNÁNDEZ-CALVENTO Jonathan LENOIR


Universidad de Las Palmas de Jules Verne University of Picardie
Gran Canaria Amiens
Canary Islands France
Spain
Samuel LETURCQ
Antonio Ignacio University of Tours
HERNÁNDEZ-CORDERO France
Universidad de Las Palmas de
Gran Canaria Karl-Johan LINDHOLM
Canary Islands Uppsala University
Spain Sweden

Pierre-Alexis HERRAULT Alberto MALTONI


University of Strasbourg University of Florence
France Italy

Michael Timm HOFFMAN Damien MARAGE


University of Cape Town Bourgogne-Franche-Comté University
South Africa Besançon
France
Hélène HOREN
Jules Verne University of Picardie Anders MÅRELL
Amiens INRAE
France UR EFNO
Nogent-sur-Vernisson
Isabelle JOUFFROY-BAPICOT France
Bourgogne-Franche-Comté University
Besançon Giorgio MARESI
France Fondazione Edmund Mach
San Michele all’Adige
Keith J. KIRBY Italy
University of Oxford
UK
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
298 Historical Ecology

Néstor MARRERO-RODRIGUEZ Giovanna PEZZI


Universidad de Las Palmas de University of Bologna
Gran Canaria Italy
Canary Islands
Spain Nicolas J. RAWLENCE
University of Otago
Justin J. MAXWELL Dunedin
Sunrise Archaeology New Zealand
Mangonui
New Zealand Vincent ROBIN
University of Lorraine
Camille MEEUSSEN Metz
Ghent University France
Belgium
Rick F. ROHDE
François MUNOZ University of Cape Town
Grenoble Alpes University South Africa
Grenoble
France Tomas ROSLIN
University of Helsinki
Claudia OLIVEIRA Finland
University of Lorraine and
Metz Swedish University of Agricultural
France Sciences
Uppsala
Bastien PARISY Sweden
University of Helsinki
Finland Ian D. ROTHERHAM
Sheffield Hallam University
Carolina PEÑA-ALONSO UK
Universidad de Las Palmas de
Gran Canaria Aarón Moisés SANTANA-CORDERO
Canary Islands Universidad de Las Palmas de
Spain Gran Canaria
Canary Islands
Emma PÉREZ-CHACÓN ESPINO Spain
Universidad de Las Palmas de
Gran Canaria David SHEEREN
Canary Islands University of Toulouse
Spain France
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
List of Authors 299

Lara D. SHEPHERD Donald M. WALLER


Museum of New Zealand Te Papa University of Wisconsin
Tongarewa Madison
Wellington USA
New Zealand
Richard WALTER
Thomas VANNESTE University of Otago
Ghent University Dunedin
Belgium New Zealand

Kris VERHEYEN Florian ZELLWEGER


Ghent University Swiss Federal Institute for Forest,
Belgium Snow and Landscape Research
Birmensdorf
Alexander J.F. VERRY Switzerland
University of Otago
Dunedin
New Zealand
and
University of Toulouse
France
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Index

A, B C, D

agrarian, 251–253, 255, cadastral maps, 47, 53


256 Castanea sativa Mill., 195
ancient DNA, 234, 241, change, 131–133
242 climate, 2, 6, 12–14, 16, 59, 61, 63
trees, 126, 130, 158 ecological, 1–4, 12
woodlands, 4, 123–125, 127, environmental, 57, 58, 65, 66, 248,
128, 131–133, 152, 154, 256
157, 158 land-use, 145
anthracology, 117 social, 199
anthropogenic, 88, 181 vegetation, 190
activities, 138, 142–144 Chatham Islands, 239, 241
archaeology, 99, 100, 145, 146, 243, chestnut civilization, 195, 196
287, 288 citizen science, 66
arctic, 222–225, 229, 230 climate warming, 260, 270
beetle, 156 co-extinction debt, 279
biodiversity, 71, 72, 77–80, 198, collaboration, 286, 288–290
202 colonization credit, 6
biological turnover, 234, 236, community composition, 224,
241 227–229
biomass harvest, 117 conservation, 124–126, 131, 133
biotic homogenization, 2, 14 baselines for, 65
birds, 233, 234, 236 nature, 152, 153, 155, 157, 158
botanical continuity, 124, 127–129, 131–133
indicators, 127, 128, 133 coppice/coppicing, 154
remains, 89, 94, 95 cultivation, 234, 237, 239, 240
survey, 64 deer, 7, 12, 13, 15
delayed extinction, 274, 277

Historical Ecology: Learning from the Past to Understand the Present and Forecast the
Future of Ecosystems, First Edition. Guillaume Decocq.
© ISTE Ltd 2022. Published by ISTE Ltd and John Wiley & Sons, Inc.
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
302 Historical Ecology

demographic stochasticity, 275 habitat


digital elevation model (DEM), 92 Directive (92/43/EEC), 198
disturbance regime, 2, 4, 5, 12 fragmentation, 2, 15
diversity, 222, 224, 225, 227 loss, 275, 279
domesticated landscapes, 210, 217 patches, 73, 77–80
quality, 274, 278, 280
E, F hay-meadows, 209, 212, 214
hedge/hedgerow, 155
ecology (see also historical
herbaceous layer, 30, 36
ecosystem), 1–7
herbivory, 166–168, 170–172
ecosystem
historic
ecology, 27
archives, 145
engineer, 164
maps, 136
historical, 274, 276, 277, 280, 281
historical ecology, 27, 45, 50, 52, 55,
restoration, 164, 172, 173
99, 196, 221, 288–291
endozoochory, 166
horizon, 86, 90, 92
environmental
human impact, 234, 238
filter, 166
hunting, 177, 184, 187, 188
history, 3
hysteretic model, 40
epizoochory, 166
Ethiopia, 58
I, L
extinction, 234, 236, 240–242
debt, 6 IHOPE, 286–291
forest, 135–141, 144–146 immigration credit, 275, 276, 280,
biodiversity, 260, 262, 266 281
canopy, 261 infield, 209, 212, 215, 216
disturbance, 118, 119 systems, 209, 212–216
edges, 262, 264, 266 integration, 287, 290
management, 116, 117, 259, 260, invasive species, 4
265, 266, 269, 270 Iron Age, 208, 210, 211, 215,
old/ancient, 281 216
REplot, 12 iron production, 207, 211–216
functional trait, 4, 12, 14 land
future, 285, 288, 290, 291 cover, 45–48, 50–53, 55
Fynbos, 59 change, 45
use, 45, 48, 247, 248, 251–253,
G, H 255
change, 259, 264
Gallo-Roman, 178, 179, 182–184,
landforms
190
aeolian, 248
geography, 99, 100
landscape archaeology, 2
geology, 99–101, 103
geomorphological processes, 88, 89
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Index 303

laser imaging detection and ranging Pinnipeds, 236


(LiDAR), 73, 142, 143, 145 plant community, 166, 171
late medieval agrarian crisis, 211 Podzol, 186
legacy, 119 pollen analysis, 142
livestock, 209, 212–215 Polynesian, 233, 235, 237, 239, 240,
local, 285, 286, 289, 290 242, 243

M, N R, S

macroclimate, 264, 265 regional, 285, 286, 288–291


management relaxation time, 275, 276
traditional, 201, 202 remote
woodland, 151, 153, 154, 158 sensing, 71, 223, 226–229, 260,
Māori, 235–240 261, 265, 267
map, 152, 153 techniques, 4
historical, 75 resurvey, 11, 16
mapping, 45, 46, 48, 50–52, 55 rewilding, 155, 158, 164, 172
microclimatic buffering, 262, 264, satellite images, 79
270 Savannah, 59
microtopography, 85, 88, 92, 95 scales, 285, 288, 290
Middle Ages, 208, 209, 211, 215 secondary succession, 136
missing baseline, 2 Shantz, H.L., 58
Moa, 235 shielings, 214, 215, 217
monitoring, 222–224, 226, 227, 229 socio-ecosystem, 3, 6
natural experiment, 18 soil, 105–107
NDVI (Normalized Difference charcoals, 186
Vegetation Index), 223, 225, 226 organic matter (SOM), 93–95
New Zealand, 233, 234, 237, 243 seed bank, 182
nitrogen South Africa, 57–59, 63, 64
deposition, 2, 5, 13, 15 southern Africa, 58–60, 66
saturation, 31, 32, 40 species
engineer, 276
O, P keystone, 278, 279
strata, 105, 106
ordination, 20
Succulent Karoo, 59, 60
organic matter (OM), 86, 87, 93
outland, 209, 212–216
T, U
P. serotina, 188, 189
palynology, 238, 240 thermophilization, 268
pedoturbation, 86 time, 99–107
phenology, 224–227, 229 timescale, 100, 102, 104–106
photographs topographic anomalies, 178
aerial, 47, 50–52, 64, 66, 75, 76 tourism, 248, 253, 255, 256
historical, 59, 60, 62, 64, 66 transhumance, 214
Downloaded from https://onlinelibrary.wiley.com/doi/ by Columbia University Libraries, Wiley Online Library on [08/07/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
productivity, 223, 225
vegetation, 221–229

written sources, 3, 4
wood-pasture, 155
V, W
transient dynamics, 273, 281

tundra, 221, 222, 224–229


ungulates, 167, 170–172
translocation, 239, 240
Historical Ecology
304

You might also like